121
UNIVERSIDADE DE SÃO PAULO INSTITUTO DE ASTRONOMIA, GEOFÍSICA E CIÊNCIAS ATMOSFÉRICAS Versão Corrigida. O original encontra-se disponível na unidadeJORGE ROSAS SANTANA Clouds and their effects on solar radiation at São Paulo (Nuvens e seus efeitos na radiação solar em São Paulo) São Paulo 2018

Clouds and their effects on solar radiation at São Paulo

  • Upload
    others

  • View
    3

  • Download
    0

Embed Size (px)

Citation preview

UNIVERSIDADE DE SÃO PAULO

INSTITUTO DE ASTRONOMIA, GEOFÍSICA E CIÊNCIAS

ATMOSFÉRICAS

“Versão Corrigida. O original encontra-se disponível na unidade”

JORGE ROSAS SANTANA

Clouds and their effects on solar radiation at

São Paulo

(Nuvens e seus efeitos na radiação solar em

São Paulo)

São Paulo

2018

JORGE ROSAS SANTANA

Clouds and their effects on solar radiation at

São Paulo

(Nuvens e seus efeitos na radiação solar em

São Paulo)

Dissertação apresentada ao Instituto de

Astronomia, Geofísica e Ciências Atmosféricas da

Universidade de São Paulo para obtenção do título de

Mestre em Ciências.

Área de concentração: Meteorologia

Orientador: Prof. Dra. Marcia Akemi Yamasoe

I

Aos meus pais

II

Agradecimentos

Queria agradecer primeiramente aos meus pais por me incentivar e

motivar no mundo da meteorologia. Ao sistema de educação de Cuba (público

e para todos) por contribuir com minha formação como meteorologista.

Agradecimento especial à minha estimada orientadora Dra Marcia Akemi

Yamasoe, por todo apoio e profissionalismo durante a pesquisa, em ensinar

novos conhecimentos que enriqueceram esta pesquisa e minha formação.

A CAPES pela concessão da bolsa de mestrado.

Aos doutores Eduardo Landulfo, Fabio Lopes por fornecer os dados do

Lidar e sky câmera.

Ao Dr. Bernard Mayer por fornecer o código LibRadtran. À AERONET,

centro de processamento de CLOUDSAT e EarthCARE Research Product

Monitor de JAXA pelos dados de aerossóis e de propriedade de nuvens de

satélite usados.

Aos doutores Edmilson Freitas e Elisa Sena por contribuírem

positivamente ao trabalho no exame de qualificação.

Ao Dr. Boris Barja Gonzalez, meu primeiro orientador, por todo seu

apoio, desde o começo, no mundo da pesquisa. Ao Dr Michel Mesquita por me

ajudar com minha permanência no Brasil.

Ao Eduardo Fernandes por ajudar com a manutenção dos equipamentos

de medição de radiação solar utilizados nesta pesquisa. Aos observadores

meteorológicos da estação meteorológica do IAG, pela sua excelente

contribuição com as observações de nuvens por quase seis décadas.

Aos trabalhadores da seção de informática do DCA do IAG.

Aos professores Dr. Rene Estevan Arredondo e Dr. Juan Carlos Antuña

com minha formação na área de transferência radiativa.

A minha querida Camila, por sempre estar do meu lado, nos bons e

maus momentos, me trazer felicidade e motivação para seguir em frente,

independente dos problemas que poderiam surgir.

Aos meus queridos amigos Luis, Talita, Veronika, Sameh, Isaque Luan,

João Basso, João Hackerott, Paola, Alberto Bié, Andrea e Miriam por

compartilharem estes maravilhosos dois anos de mestrado. Aos meus

III

companheiros cubanos, Darsys & Raidiel, Janet & Jose, Yusvelis, Maciel,

Damian & Jenifer e Dayana por me apoiarem no começo do mestrado.

IV

Resumo Rosas J. Nuvens e seus efeitos na radiação solar em São Paulo, 2018. Dissertação (Mestrado). Instituto de Astronomia, Geofísica e Ciências Atmosféricas, Universidade de São Paulo, São Paulo, 2018.

Na Região Metropolitana de São Paulo, foram estudadas as nuvens e

seus efeitos na radiação solar. Para tanto, foram usadas observações visuais

de nuvens, medições desde a superfície efetuadas por diferentes radiômetros,

produtos dos satélites de órbita polar CALIPSO e CloudSat e o modelo de

transferência radiativa 1-D LibRadtran.

Foi desenvolvida uma climatologia para o ciclo diurno da fração de

cobertura de nuvens (1958-2016) usando dados de observações visuais. O

ciclo diurno da cobertura de nuvens foi dominado por nuvens baixas,

especialmente as estratiformes. Observaram-se diferenças entre o ciclo diurno

das nuvens baixas cumuliformes e estratiformes. Além disso, houve uma

tendência de aumento da fração de cobertura de nuvens baixas (1,6

%/década), especificamente das estratiformes (3,1 %/década), e das nuvens

cirriformes (0,8%/década). Por outro lado, observou-se tendência de diminuição

da fração de cobertura de nuvens médias (-2,4%/década).

A variabilidade sazonal e diurna do perfil vertical de nuvens foi

analisada, com as nuvens atingindo maiores altitudes à noite e no verão. No

inverno, as nuvens baixas predominaram.

A profundidade óptica efetiva da nuvem (ECOD), usando a transmitância

total em 415 nm, e os efeitos instantâneos das nuvens sobre a radiação solar,

de medições de irradiância solar global, foram estimados em sinergia com

cálculos feitos com o LibRadtran. ECOD apresentou variabilidade diurna e

sazonal com máximo na primavera (34,4) e no período da tarde (34,2) e

mínimo pela manhã, próximo ao nascer do sol (25,5) e no inverno (26,9) para

nuvens baixas. O efeito radiativo de onda curta apresentou dependência com

relação à obstrução do disco solar pelas nuvens, o tipo de nuvem e fração de

cobertura. A atenuação máxima foi observada para nuvens baixas com o céu

totalmente nublado, com valor médio de redução de 72 % da irradiância global,

comparada com condições de céu claro. Medianas de redução de nuvens

médias e altas foram de 57 % e 33 %, respectivamente. Foram observados

V

efeitos de incrementos da radiação solar (enhancement) de cerca de 10 %

com duração de até 20 minutos, devido ao espalhamento pelas laterais das

nuvens, em presença de todos os tipos de nuvens analisados, quando o disco

solar não estava obstruído. O máximo de enhancement chegou até 50 % na

presença de nuvens baixas.

Palavras-chave: Nuvens, COD, MFRSR, CALIPSO-CloudSat,

LibRadtran.

VI

Abstract

Rosas J. Clouds and their effects on solar radiation in São Paulo, 2018. Dissertation (Master)-Instituto de Astronomia, Geofísica e Ciências Atmosféricas, Universidade de São Paulo, São Paulo, 2018.

Clouds and their instantaneous effects on downward solar radiation were

studied at the Metropolitan Area of São Paulo. For this purpose, visual

observations of clouds, ground-based measurements performed by different

radiometers, products from the polar orbiting satellites CALIPSO and CloudSat

and 1-D Radiative Transfer Model (RTM) LibRadtran were used.

Daytime climatology of cloud cover fraction (1958-2016) using data of

hourly visual observations was carried out. The diurnal cycle of cloud cover

fraction was dominated by low clouds especially by stratiform clouds.

Remarkable differences in the diurnal cycles of low cumuliform and stratiform

clouds were also observed. During the time period, positive trends for low cloud

cover (1.6 %/decade), especially stratiform (3.1 %/decade), and cirriform cloud

(0.8 %/decade) were observed, while a decreasing trend of mid-level cloud

cover (-2.4%/decade) was found.

Seasonal and diurnal variability of vertical profile of cloud was observed,

with cloud extending to higher altitudes at night and with maximum frequency of

occurrence observed in summer. In winter, low clouds prevailed.

Effective cloud optical depth (ECOD), using the total transmittance at

415 nm, and instantaneous cloud effects on solar radiation at the surface, using

global irradiance measurements, were estimated in synergy with LibRadtran

computations. ECOD presented seasonal and diurnal variability, with maximum

of mean in spring (34.4) and in the afternoon (34.2), and minimum at sunrise

(25.5) and winter (26.9) for low clouds. The shortwave effects of clouds

depended on solar disk condition, cloud type and cloud cover. Maximum of

shortwave radiative attenuation was observed for low clouds in total overcast

conditions with a median reduction of 72 % of global irradiance compared to

clear sky. Median reduction of mid and high clouds was 57 % and 33 %,

respectively. Enhancement effects with duration as long as 20 minutes, caused

by lateral scattering, were observed in the presence of all analyzed cloud types,

VII

when the solar disk was not blocked by clouds, increasing global solar

irradiance around 10% at the surface. Maximum enhancement could reach 50

% for low clouds.

Key words: clouds, COD, MFRSR, CALIPSO-CloudSat, LibRadtran.

VIII

Figures Index

Figure 1: Examples of idealized shapes of the particle habits. Clockwise from top left: dendrite, aggregate, bullet rosette, solid column, hollow column, and plate (from Key et al., 2002). ..................................................................... 11

Figure 2: Asymmetry parameter (g) (a), single scattering albedo (𝝎) and the ratio between the extinction coefficient and LWC or IWC (𝛽𝑒 /LWC or IWC) for liquid and ice clouds. For liquid clouds, Hu and Stamnes (1993) parameterization is employed and for ice clouds, the parameterization of Key et al. (2002) for rough-aggregate habit is used. The optical properties are computed for re values of 5,10 and 15 µm for liquid clouds and for 30, 51.2 and 70 µm for ice clouds. ........................................................................................ 20

Figure 3: Mean spectral surface albedo used as input in the uvspec radiative transfer model. ................................................................................... 41

Figure 4: Mean profiles of normalized aerosol extinction coefficient for each season and time of the day (a) and the normalized profile of AOD built by synergy between LIDAR and sun-photometer for normal or typical conditions and for conditions with biomass burning plume advection in spring (SON) (b). 42

Figure 5: Flow chart of the main procedures carried out for the ECOD retrieval and computations of cloud radiative effects. ....................................... 45

Figure 6: Variations of diffuse ratio, at channels 870 nm and 415 nm, at surface level, with CSZA and AOD for clear sky, and COD of ice and liquid clouds computed from 1-D uvspec model of LibRadtran. The darker the color the higher the value of AOD500 or COD. COD for ice clouds spans from 0.03 up to 4 and, for liquid clouds, it varies between 3 and 100. AOD500 is varied between 0.08 and 0.8. Between 0.08 and 0.3 the step is 0.03, while above 0.3 is 0.1. ................................................................................................................ 48

Figure 7: Profiles of frequency of occurrence of clouds per season at nighttime (a) and daytime (b), computed from cloud mask generated by merged CALIPSO-CloudSat data. ................................................................................. 54

Figure 8: Frequency of presence of clouds relative to the distance from the coast line. ................................................................................................... 55

Figure 9: Seasonal variations of geometrical properties of clouds: the height of base and thickness of liquid clouds (a, b) and ice clouds (c, d). The number of clouds identified is also shown in the upper border of each figure. . 56

Figure 10: Microphysics properties re and LWC or IWC for liquid (a, b) and ice clouds (c, d). For liquid clouds, the products of CloudSat are employed and for ice cloud, CSCA-Micro product was considered. The number of cases is placed in the upper border of each figure. ........................................................ 57

Figure 11: Profiles of frequency of occurrence of ice cloud habit for each season (a) and frequency of occurrence of cloud layer totally composed by a given cloud particle habit. ................................................................................. 58

Figure 12: Variations of re and IWC with height above cloud base (a, b) and variations of mean re and IWC of the cloud layer according to the height of cloud base (c,d). ............................................................................................... 59

IX

Figure 13: Mean diurnal cycle and standard deviation (a) computed from mean diurnal cycles calculated for every 15 days. Variability of the diurnal cycle for every 15 days computed as hour of maximum (b) and Amplitude (c). ........ 60

Figure 14: Mean annual cycle computed for every 15 days. The vertical bars represent the inter-annual variability. ....................................................... 62

Figure 15: Mean values of cloud cover (amt) per year computed for every cloud type in summer (a,b ), autumn (c,d ), winter (e,f), spring (g,h). ............... 64

Figure 16: Mean frequency of occurrence (fq) and sky cover when present (awp) for the12 main cloud combinations observed at each season. .. 69

Figure 17: Diurnal cycles of the 8 main combinations of clouds for frequency of occurrence (a) and amount when present (b). ............................. 70

Figure 18: Sensitivity of transmittance at 415 nm for liquid and ice clouds to variations of aerosol vertical profile [mo_ty= morning typical, mo_sp= morning in spring, af_ty=afternoon typical, af_sp=afternoon in spring, su-sp=summer-spring and au-wi=autumn-winter given by LibRadtran (Mayer et al., 2015) (a), variations of AOD415 (b), variations of SSA440 (c) and g440 (d). ................... 71

Figure 19. As in the Figure 16 but for CSZA (a), surface albedo at 415 nm (b), dioxide of nitrogen (NO2) (c) and ozone (O3) (d). ................................ 72

Figure 20. Variations of total transmittance at 415 nm to cloud base, thickness, re and LWC for ice and liquid clouds with a constant value of COD. 72

Figure 21. Relative error of COD retrieval due to errors in each variable: AOD, SSA, T, re and CB and the total error for each COD value. For low clouds (a) and high clouds (b). .................................................................................... 73

Figure 22: Total relative error of COD retrieval, for different methods of assuming AOD in cloudy cases, for low clouds (a) and ice clouds (b). ............ 75

Figure 23: Instantaneous global irradiance modeled by 1-D uvspec and measured by pyranometer (a), frequency distribution of absolute differences between modeled and measured values (b), and relative differences using no interpolated AOD and water vapor (not interpolated) (c), interpolated on the day or adjusted from monthly diurnal cycle (interpolated_1) (d) and AOD and water vapor computed from mean monthly diurnal cycles (interpolated_2) (e). ......... 76

Figure 24: Relative differences of T at 415 nm modeled versus T at 415

nm measured by MFRSR 345 (a,c) and of 𝑇𝑑𝑖𝑟 modeled and measured at 415 nm (b,d). ........................................................................................................... 77

Figure 25: Validation of ratio of downwelling diffuse irradiance (D↓) at 870 nm and 415 nm (DR) for cases discriminated by the all-sky camera (a), the classification defined from model and observed values (c), for a specific day with low broken clouds (b) and for a specific day of low clouds with total overcast condition (d). ...................................................................................... 78

Figure 26: Comparison of modeled global irradiances using percentile 50 (a,d,g), percentile 5 (b,e,h) and percentile 95 of re for ECOD values retrieved for low clouds with total overcast condition. ........................................................... 80

Figure 27: As Figure 25 but for cirrus clouds with total overcast. ........... 81

X

Figure 28: Comparison between COD retrieved from AERONET and the COD retrieved by MFRSR under the presence of low clouds in total overcast condition. Time series for one specific day (a), COD of AERONET vs COD of MFRSR for coincident cases (b). G modeled using COD AERONET vs G measured by pyranometer (c). G modeled using COD from MFRSR vs G measured by pyranometer (d). Differences with respect to CSZA for AERONET (e) and MFSRSR (f). Frequency distribution of absolute error for AERONET (g) and MFRSR (h). ............................................................................................... 82

Figure 29. Seasonal and hourly COD distribution for low clouds. For summer (a), autumn (b), winter(c), spring (d). From sunrise up to 08:00 LT (e), between 08:00 and 10:00 LT (f), around noon (from 10:00 up to 14:00 LT) (g) and after 14:00 LT(h). ....................................................................................... 84

Figure 30. As Figure 28 but for cirrus clouds. ........................................ 84

Figure 31: Cloud effects computed for cloudiness conditions and solar disk. Subscript 0 indicate obstructed solar disk and 1 not obstructed by clouds. Cloudiness are LW=low clouds, LWBK=low broken clouds, H=High clouds with total overcast, HBK=High broken clouds, MID=mid-level clouds with total overcast, MID_BK =mid-level broken clouds. ................................................... 86

Figure 32: Specific day with clear sky conditions (a), low broken clouds (b,d) and total overcast (c)................................................................................ 87

Figure 33. Relationship between CRE vs COD for liquid clouds (a) and ice clouds (c). Relationship between NCRE and COD for liquid clouds (b) and ice clouds (d). Colors are related to central values of CSZA. Value of m is the slope and R2 the determination coefficient of the curve Y =m lnCOD + n. ....... 89

Figure 34. Cloud efficiency computed for CRE (a) and NCRE (b) for different CSZA values as function of COD for liquid clouds. Comparison between CEF (from NCRE) vs COD for liquid and ice clouds, for CSZA centered at 0.85 (c). Seasonal variation of CEF (from NCRE) for liquid clouds (d). .................................................................................................................... 90

XI

Tables Index

Table 1: The 10 types of cloud genera as classified by WMO, according to the height level. .............................................................................................. 6

Table 2: Default cloud configuration assumed in the radiative transfer code. ................................................................................................................ 44

Table 3: Sky definition algorithm for the clouds scenarios defined in this research. .......................................................................................................... 49

Table 4: Seasonal and annual trends (%/decade) computed for each cloud type, in three different time intervals. Statistical significance was computed by the modified Mann-Kendall Test at 5 % of confidence level and they are highlighted in bold............................................................................... 64

Table 5: Trends (%/decade) of 6 six clouds types computed for sunrise (7 SLH), noon (13 SLH) and sunset (18 SLH). Statistical significance was computed by Mann- Kendall Test in 5 % and they are in bold.......................... 66

Table 6: Frequency of occurrence and co-occurrence for clouds types for each season ..................................................................................................... 67

Table 7: Error of estimating AOD440 for cloudy conditions of each method: time interpolation throughout the day (Inter 1), monthly mean diurnal cycle of AOD (Inter 2), adjusted monthly mean diurnal cycle from the AOD measured in the previous time (Inter 3) and later time (Inter 4), adjustment of the climatological monthly diurnal cycle of AOD from mean diurnal AOD value computed as close as at least 2 days (Inter 5). ................................................ 75

XII

Acronyms and abbreviations

AERONET AErosol RObotic NEtwork

AOD Aerosol Optical Depth

CALIPSO Cloud-Aerosol Lidar and Infrared Pathfinder Satellite

Observations

CB Cloud base height

CEF Shortwave Cloud Efficiency

CloudSat Cloud Satellite

COD Cloud Optical Depth

CRE Shortwave Cloud Radiative Effect

CSZA Cosine of solar zenith angle

D Diffuse Irradiance

ECOD Effective cloud optical depth

g Asymmetry Parameter

G Global Irradiance

IWC Ice Water Content

LWC Liquid Water Content

MASP Metropolitan Area of São Paulo

MFRSR Multifilter Rotating Shadowband Radiometer

NCRE Shortwave Normalized Cloud Radiative Effect

NIR Near Infrared

re Effective radius

R Hemispherical Reflectance

XIII

RTE Radiative Transfer Equation

S Direct Sun Radiance

SS Direct Sun Irradiance

SS0 Direct Sun Irradiance at Top of the Atmosphere

SSA Single Scattering Albedo

SZA Solar zenith angle

T Total hemispheric transmittance

Tdir Transmittance of direct beam radiation

TOA Top of the Atmosphere

VIS Visible

WMO World Meteorological Organization

XIV

Contents

Figures Index ........................................................................................ VIII

Tables Index ........................................................................................... XI

Acronyms and abbreviations .................................................................. XII

1. Introduction .......................................................................................... 1

1.1. Aims .............................................................................................. 4

2. Review of literature .............................................................................. 6

2.1. About clouds .................................................................................. 6

2.1.1. Cloud types ............................................................................. 6

2.1.2. Cloud formation ....................................................................... 8

2.2. Aerosol effects on clouds............................................................. 11

2.3. Sao Paulo weather ...................................................................... 12

2.4. Shortwave radiative transfer ........................................................ 13

2.4.1. Cloud optical and radiative properties ................................... 18

2.4.2. Retrieving COD and re from ground based measurements ... 22

2.4.3. Computing the shortwave radiative effects of clouds ............ 25

2.4.4. Cloud radiative efficiency (CEF) ............................................ 26

2.4.5. The enhancement effect of clouds on the solar radiation ...... 27

2.4.6. The shortwave attenuation effect of clouds ........................... 28

3. Instruments and methods .................................................................. 30

3.1. Satellite data ................................................................................ 30

3.1.1. Description of satellites ......................................................... 30

3.1.2. CloudSat and CALIPSO products ......................................... 31

3.1.2.1. CloudSat ............................................................................ 31

3.1.2.2. CloudSat-CALIPSO (JAXA products) ................................. 32

3.2. Ground based measurements ..................................................... 34

3.2.1. Visual observations of clouds ................................................ 34

XV

3.2.2. Pyranometer .......................................................................... 34

3.2.3. Multifilter Rotating Shadowband Radiometer (MFRSR) ........ 35

3.2.4. AERONET sun-photometer ................................................... 35

3.2.5. Sky camera ........................................................................... 36

3.2.6. Lidar ...................................................................................... 37

3.3. Methodology of cloud climatology from visual observation .......... 37

3.4. Radiative transfer model (uvspec from LibRadtran) .................... 39

3.4.1. Surface albedo ...................................................................... 40

3.4.2. Aerosols ................................................................................ 41

3.4.3. Clouds ................................................................................... 43

3.5. Methodology of the radiative computations and COD retrieval. ... 44

3.5.1. Sky definition ......................................................................... 45

3.5.1.1. Sky definition by visual observations .................................. 45

3.5.1.2. Direct sun criterion (solar disk obstruction) ........................ 46

3.5.1.3. Spectral diffuse ratio (DR) .................................................. 47

3.5.2. Calibration of MFRSR 368 .................................................... 49

3.5.3. The retrieval of effective cloud optical depth ......................... 50

3.5.3.1. Error in COD retrieval ......................................................... 51

3.5.4. Instantaneous cloud effects on solar radiation ...................... 52

4. Results ............................................................................................... 54

4.1. Cloud characteristics from CALIPSO-CloudSat ........................... 54

4.2. Cloud climatology from visual observations ................................. 59

4.2.1. Diurnal and annual cycles of cloud amount ........................... 59

4.2.2. Inter-annual variability of cloud amount ................................. 63

4.2.3. Simultaneous occurrence and combinations of clouds .......... 66

4.3. Sensitive study for COD retrievals ............................................... 70

4.3.1. Sensitive of properties at 415 nm .......................................... 70

XVI

4.3.2. Errors affecting COD retrieval ............................................... 73

4.3.2.1. Error of AOD estimation into COD retrieval ........................ 74

4.3.3. Model response to clear sky global irradiance ...................... 75

4.3.4. Spectral diffuse ratio validation.............................................. 77

4.4. Effective cloud optical depth (ECOD) .......................................... 79

4.4.1. Validation with pyranometer ground-based measurements .. 79

4.4.2. Comparison with COD retrieved from sun-photometer .......... 80

4.4.3. Effective cloud optical depth, seasonal and diurnal variability.

................................................................................................................... 83

4.5. Cloud effects on solar radiation ................................................... 85

4.5.1. Shortwave cloud radiative effects .......................................... 85

4.5.2. Cloud efficiency ..................................................................... 88

5. Conclusions ....................................................................................... 92

6. Future works ...................................................................................... 95

7. References ........................................................................................ 96

1

1. Introduction

Clouds are the main component of the radiative transfer at the

atmosphere. They absorb and emit terrestrial radiation, warming the earth

surface and the atmospheric layers underneath (positive effect) and cooling the

atmospheric layers aloft. These processes also contribute to the formation and

development of clouds (Heymsfield et al., 2017; Wood, 2012).

In the shortwave spectrum, clouds have an elevated albedo, producing a

cooling radiative effect (negative effect) on beneath layers and the surface. The

magnitude and signal of the radiative effect depends on the properties of the

clouds. The main shortwave effect of low clouds at surface is cooling, but there

are cases with short durations, when they do not block the solar disk, that

warming effects or enhancement of solar radiation at the surface can occur

(Mateos et al., 2013; Tzoumanikas et al., 2016). Cirrus clouds play a different

role compared to other clouds, they warm the atmosphere more effectively,

because they are optically thin and located at higher altitudes (Stephens, 2005).

Clouds are also important to the hydrological cycle of the earth, bringing

water to the surface by precipitation and transporting water vapor to higher

layers of the troposphere. The condensation process releases a large amount

of latent heat, contributing to the transport of energy in the atmosphere

(Boucher et al., 2013). All these processes related to clouds affect large-scale

circulations and wave disturbances. Hence, cloud systems are important issues

for the climate and weather prediction numerical models.

Clouds can respond to the global warming due to the increase of CO2

concentration, a process known as cloud feedback. Therefore they can

contribute to mitigate or reinforce the warming effect (Lee, 2016). Theoretically,

the global warming can strengthen due to cloud positive feedback: increasing

the tops of high clouds by expansion of the troposphere and the expansion of

the Hadley cell, producing the migration of storm tracks poleward. The latter can

increase low clouds near poles, where solar radiation at the surface is reduced,

decreasing the shortwave radiative effect (Boucher et al., 2013). Cloud cover

fraction of low, mid and high clouds is expected to decrease specially in

subtropical areas, but that is still unclear (Boucher et al., 2013).

2

Owing to cloud positive feedback, global observations and simulations

indicate that cloud properties such as cloud amount are changing over the time.

Eastman and Warren, (2013) reported the decrease of mid and high level

clouds in mid latitudes with a global decrease of cloud amount of about 0.4 %

per decade. Recently, Norris et al. (2016) showed agreement between satellite

records and climate simulations of the reduction of cloud cover globally. They

observed an increase of the height of higher clouds in all latitudes and the

expansion of the subtropical dry zones as a consequence of the increase of

greenhouse gas concentration and a recovery from the volcanic eruptions.

From the above mentioned reasons, understanding the role of clouds in

the climate system is needed. Studies focused on the possible changes of the

diurnal cycle and regional cloud amount are also crucial (Eastman and Warren,

2014). Global circulation models have large bias in cloud frequency occurrence

of cloud types and their respective diurnal cycles. In addition, there is a need of

measurements of cloud properties and studies of their radiative impacts in

climate-sensitive areas (Burleyson et al., 2015).

Nowadays, with the inclusion of active sensors as radar and lidar on

satellites, these techniques become strong tools for studying the vertical profiles

of clouds. This is important to understand the impact of clouds on the radiation

budget of the earth (Joiner et al., 2010) and the impact of the vertical

distribution of latent heat released by them on the global circulation and

precipitation. The A-train constellation formed by Cloud Satellite (CloudSat) and

Cloud-Aerosol Lidar and Infrared Pathfinder Satellite Observations (CALIPSO),

among other satellites, have contributed to the improving of retrieving cloud

properties and their associated radiative forcing. In addition, the observations of

clouds from the surface still represent a useful dataset with some advantages

over satellite records. First, a long period of records, duration and temporal

resolution allows for a better study of trends and diurnal cycles (Eastman and

Warren, 2014). Besides, according to the authors, ground based observations

are still the best detector of low clouds.

Studying the clouds over urban and industrialized areas is a focus of high

interest, because changes of their properties can be induced by the increase of

3

aerosol concentrations, as a consequence of air pollution and by aircraft

emissions. Furthermore, cloud properties can change due to the modification of

land use, which, in turn, can alter the hydrological cycle. Cloud albedo can

increase and cloud can last longer in the atmosphere, contributing to the higher

shortwave cooling effect. However, cloud cover can be reduced, by the increase

of aerosols with high absorption efficiency of solar radiation, reducing the

shortwave cooling effect. There are contrasting results about how aerosols

affect the evolving clouds (Boucher et al., 2013), therefore it is still a topic where

many studies are needed.

An important example of region with higher urbanization and widely

extended area is the megacity of São Paulo, considered as the top ten of larger

cities around the world. It is an important source of aerosols generated by

emission of cars and also influenced by long range transport of biomass burning

plumes in spring (de Almeida Castanho et al., 2008). The precipitation regime is

changing, with extreme events being more frequent. Especially since the end of

the 1950s, a significant increase of rainy days and total daily rainfall have been

observed (Obregón et al., 2014; Silva Dias et al., 2013). Silva Dias et al. (2013)

also mentioned that the intensification of the urban heat island and cloud

microphysics modification due to pollution are factors to take into account for

explaining the positive trend observed in the wet season in the last eighty years.

The study of clouds in the region is an issue to be addressed due to the

possible influence of urbanization on their properties (Yamasoe et al., 2017).

At São Paulo, few works related to clouds and their impacts on solar

radiation have been carried out. They are focused on the specific topics:

Moura et al. (2016) studied total cloud cover variability at São Paulo from

visual observations during 1961-2013. The work is limited to the mean total

cloud cover in terms of the diurnal and annual cycles. They compared mean

diurnal and annual cloud cover from visual observations with cloud parameters

retrieved from satellite and found good agreement.

Landulfo et al. (2009) showed the ability to obtain optical properties of

thick cirrus clouds in the region from ground-based LIDAR at 532 nm and 355

nm. They discussed the frequency of occurrence of cirrus as computed for valid

4

days of measurements when low clouds did not affect the LIDAR beam. Due to

the restricted field of view of the system, the measurements lead to unrealistic

results of frequency of cirrus clouds, for instance, maxima observed in winter.

Recently, Yamasoe et al. (2017) studied the climatological radiative

effect of low, multilayer and high level clouds based on 24 h mean of downward

solar irradiation data at the surface. They analyzed the seasonal cycle of cloud

cover and how the solar radiative effect changed with the sky fraction covered

with clouds with distinct base height. Considerable differences in cloud effects

between January and July were observed, due to the high dependence on the

cosine of the solar zenith angle (CSZA). The work is limited, since the authors

based their analysis on the cloud fraction only, without discrimination of cloud

genera, and did not compute the instantaneous cloud radiative effect on solar

radiation, which implies the knowledge of other cloud parameters, such as cloud

optical depth. Thus, trying to complement previous studies on the effect of

clouds on solar radiation at São Paulo region, the aims of this work are

presented next.

1.1. Aims

This project aims the characterization of clouds in the Metropolitan area

of Sao Paulo (MASP) and the estimation of their effects on solar radiation

reaching the surface. The specific objectives are:

1. Understand how the physical and optical properties vary according to

cloud types.

2. Analyze long term variability of cloud cover fraction.

3. Estimate the shortwave radiative effects of clouds using the radiative

transfer library LibRadtran.

The objectives will be carried out with the next tasks:

Analysis of clouds and their properties using products generated by the

polar orbiting satellites CloudSat and CALIPSO.

Retrieve optical properties of clouds from ground-based measurements

of radiative fluxes.

5

Long term clouds climatology using visual observations of cloud cover

fraction, according to cloud genera and base height.

Shortwave cloud radiative effects and efficiency computed from cloud

properties obtained from ground-based measurements and the

LibRadtran.

6

2. Review of literature

2.1. About clouds

2.1.1. Cloud types

The World Meteorological Organization (WMO) defines 10 types of

clouds ‘genera’ according to the height level where they form and to their

appearance, as presented in Table 1. These 10 genera make a total of 100

combinations of species and varieties, describing the shape, internal structure,

transparency and arrangement of clouds (https://cloudatlas.wmo.int/). Prefixes

and suffices from 'Latin' indicate the character of clouds - stratus: flat; cumulus:

heap; cirrus: feathers, wispy; nimbus: rain; alto: mid-level.

Table 1: The 10 types of cloud genera as classified by WMO, according to the height level.

Level Height Acronym Name

High Above 6000 m Ci Cirrus Cc Cirrostratus Cs Cirrocumulus

Mid 2000-8000 m As Altostratus Ac Altocumulus Ns Nimbostratus

Low Below 2000 m Cu Cumulus Sc Stratocumulus St Stratus Cb Cumulonimbus

Stratocumulus (Sc), the combination of Latin stratus and cumulus, is a

low cloud type as a result of grouping convective elements forming a layer. It is

the most frequent cloud coverage around the world. Annually about 23 % of the

sky over the ocean, with maximum of 60 % in the colder regions of sea

anticyclone systems (Wood, 2012), are covered by Sc clouds. Over the land,

the mean fraction of the sky covered by Sc is around 12 % (Warren et al.,

2007). The seasonal cycles of Sc in the southern hemisphere are stronger than

in the northern hemisphere and the month of maximum cloud cover is variable

and depends on the region. Although the influence of the seasonality is unclear

in the tropical regions, at the North Atlantic and Pacific Oceans, the maximum

Sc cloud cover is observed in winter, in the western sides and, in summertime,

in the eastern sides (Wood, 2012). The Sc has maximum coverage in the

7

morning with the higher amplitude in the eastern ocean (Eastman and Warren,

2014).

Cirrus clouds are the thinner clouds usually semitransparent to solar

radiation unless the sun is near the horizon. Cirrus can be dense enough to

produce shadows in the case of cirrus spissatus. Cirrus (Ci) and cirrostratus

(Cs) are primarily composed by ice crystals. Cirrocumulus (Cc) is highly or

totally composed of supercooled water droplets. They are inconsistent, grainy

and very thin with thickness less than 200 m (Holton and Curry, 2003). Cirrus

clouds (Ci, Cc, Cs) grouped into a single class is the most important cloud by

spatial coverage over land with 22 %, while over the sea, the mean cloud cover

is only 12 % (Warren et al., 2007). The higher frequency of cirrus clouds is

observed in the tropics near the Inter-Tropical Convergence Zone (ITCZ)

especially over the land with maximum of 70 % in the West Central Pacific

(Heymsfield et al., 2017).

Nimbostratus (Ns) are classified as mid-level clouds, but sometimes can

be observed at low levels. They produce moderately steady precipitation often

during hours. They cover only about 5 % around the globe in land and ocean

and can be in companion of convective clouds as cumulonimbus (Cb). Ns is

more frequent, with greater spatial coverage in mid-latitudes and polar regions

(https://atmos.washington.edu/CloudMap). Values of mean cloud coverage can

be as high as 24 % in polar regions.

The altocumulus clouds (Ac) look similar to Sc, but with base at mid-level

altitude and predominantly composed by water and supercooled water droplets.

Even though they produce gray shading, they are relative thin with maximum

thickness up to 1 km (Holton and Curry, 2003). They are the most important

mid-level clouds with a mean coverage of 17 % (Warren et al., 2007).

Altostratus clouds (As) are dominated by ice particles giving a diffuse and

fibrous look. They are deeper than Ac with depth values higher than 2 km, with

top usually compared to cirrus clouds. Despite the presence of As implies larger

spatial coverage, they are less frequent than Ac, leading to global mean cloud

cover of 5 % (Warren et al., 2007).

8

Cumulus (Cu) belongs to the convective clouds at the first stage. When

convection is strong enough they can develop into Cumulonimbus clouds (Cb).

Before becoming Cb, Cu can pass through a variety of steps from cumulus

fractus up to cumulus congestus. Cu has less spatial coverage as compared to

low stratiform clouds as stratus (St) and Sc. They have a global mean cloud

cover in ocean (13 %) higher than over land (5 %) (Warren et al., 2007).

Concerning the diurnal cycle, Cu presents maximum of spatial coverage near

midday over the land, with the highest amplitude over land due to the lower

thermal inertia as compared to the ocean (Eastman and Warren, 2014).

Cb is the less frequent cloud genus with mean cloud cover of 5 %, with

few differences between ocean (6 %) and land (4 %) (Warren et al., 2007).

Notable contrast is found between the diurnal cycles in ocean and land. Over

land, the maximum is observed near sunset with higher amplitude, while over

the ocean it is observed near sunrise (Eastman and Warren, 2014).

2.1.2. Cloud formation

Clouds can form when the air becomes supersaturated with respect to

ice or water, i. e. the vapor pressure of the air parcel is higher than the

saturated vapor pressure over the flat surface of the water or saturated vapor

pressure over a flat surface of ice. The presence of aerosols, acting as cloud

condensation nuclei (CCN) or ice nuclei (IN), contributes to break the energy

generated by surface tension of the drop, decreasing the equilibrium vapor

pressure over the droplet and favoring the formation and growth of cloud

particles under low values of supersaturation. The level of supersaturation

affects the size of particles activated, i.e. with the increase of supersaturation,

more smaller particles are activated (Yau and Rogers, 1989). The

supersaturation increases with the updrafts of clouds, therefore with higher

updrafts more smaller drops are expected to be activated (Stith et al., 2002).

Aerosol chemical composition and size also influence the activation of cloud

droplets, the bigger, more hygroscopic, is the aerosol the lower supersaturation

needed to activate cloud droplets (Yau and Rogers, 1989).

The air becomes supersaturated by means of two processes: the

decrease of temperature under constant pressure or advection of humid air to

9

the region contributing to the increase of the water vapor concentration.

Additionally, moist air can be forced to lifting and, in this ascension, the

temperature decreases by adiabatic expansion, vapor pressure rises and air

becomes saturated.

Ice nucleation, i.e., formation of ice by freezing, can be heterogeneous or

homogenous. The homogenous nucleation is a spontaneous freezing of super

cooled droplets or a pure solution of aerosol particles (Heymsfield et al., 2017).

It occurs at temperature lower than -38 °C and higher supersaturation on ice

(above 140 %). The heterogeneous nucleation is the most frequent mechanism

of ice formation but still not well understood because it can occur under several

conditions (Heymsfield et al., 2017). Basically, heterogeneous nucleation takes

place when temperature is below 0 °C and the air saturated with respect to ice.

It needs an IN immersed or in contact to super cooled water droplets or for

deposition of water vapor onto it. The size of supercooled drops influences the

temperature at which nucleation occurs, the larger the drops the higher the

temperature for the ice nucleation (Yau and Rogers, 1989).

There is a variety of mechanisms in the atmosphere leading to the

formation of different cloud types. Convection is the main driver of cloud

formation. By this mechanism, the air is transported to the higher layers of the

atmosphere. When convection is deeper, updrafts are strong enough to develop

deep convective clouds. The own process of convective activity can lead to the

formation of adjacent clouds such as Ns, Sc, and Cu, as well as mid and high

level clouds (Houze, 2014).

Low stratiform clouds, mid and high clouds can develop convective

instability driven by the longwave radiative cooling. The emission of longwave

radiation by the cloud layer increases the cooling at cloud top, while absorption

increases and the warming near the cloud base. These temperature differences

enhance the convective circulation previously formed by buoyancy and wind,

allowing the increase of cloud lifetime (Houze, 2014).

Low stratiform clouds of the boundary layer are coupled to the moist

source at the surface. Stratus clouds can form when they separate from the fog

layer at the surface that deepens due to mixing and radiative cooling at the top.

10

At sunrise, with solar radiation, the layers of clouds warm and hence tend to

disappear (Houze, 2014). Sc also forms under stable conditions of the lower

troposphere, when the layer of clouds is initially formed and the radiative

cooling increases the inversion layer of temperature above cloud top. This

process starts the convection that enhances as a result of the heating by

condensation in updraft and cooling by evaporation in downdraft. When solar

heating is present, the convection and layer of temperature inversion tend to

disappear and hence the own cloud (Wood, 2012).

Ns are formed by motions of near stable air that deepens enough to form

particles of rain or snowflakes. They are related to synoptic systems with

organized storms as fronts, tropical cyclones and mesoscale convective

systems (Houze, 2014).

Cirrus clouds form primarily in the upper troposphere where

temperatures are lower than -30 °C, about 8 km. Below -38 °C there is no

activation of liquid water drops, because water vapor can only be saturated or

supersaturated with respect to ice. Then, cirrus can form by lifting of moist air or

liquid or mixed phase clouds previously formed below. Mixed phase clouds are

formed by strong updrafts (Heymsfield et al., 2017).

Lifting can occur on large scales, along a frontal boundary or by small

scale vertical circulation developed around a core of jet stream. Cirrus can form

under heterogeneous or homogenous nucleation and, depending on the

mechanism, differences in ice sizes and concentrations are found (Heymsfield

et al., 2017). Radiative cooling at cloud top and radiative warming near cloud

base due to longwave emission and absorption, respectively, lead to the

development of cloud top and turbulence that contributes to maintain and

enhance the cirrus layer (Heymsfield et al., 2017).

Cirrus ice crystals have not simple idealized hexagonal shapes, instead a

variety of geometries (habit) are found. For radiative purposes ice habits are

classified as: solid column and hollow column, plate, dendrite, aggregate, bullet

rosettes (Figure 1) (Key et al., 2002).

11

Figure 1: Examples of idealized shapes of the particle habits. Clockwise from top left: dendrite, aggregate, bullet rosette, solid column, hollow column, and plate (from Key et al., 2002).

An association between ice cloud habit with temperature and humidity

has been previously reported. For stronger cloud updrafts , hexagonal and 3-D

crystals are expected, whereas for weaker velocities the plates are common

(Heymsfield et al., 2017). For temperature ranges between -20 °C and -40 °C

with low supersaturation, plates are present. Below -40 °C there is an increase

of columnar shapes. In the same range of temperature but with supersaturation

above 25 % bullet rosettes, long columns containing poly-crystals are observed

(Heymsfield et al., 2017). With cooler conditions above -60 °C, needle and

columnar forms are frequent. Stith et al. (2002) suggested aggregates as the

mean habit of ice particle in the upper regions of tropical convective clouds and

the increase with altitude of particle sizes due to aggregation effects.

2.2. Aerosol effects on clouds

Tropospheric aerosols can affect the clouds by cloud-aerosol interactions

known as Twomey effect (Twomey, 1977) and due to the rapid adjustments as

a consequence of the aerosol direct radiative effect on the surface energy

budget, the atmospheric thermodynamic profile and on the clouds (Boucher et

al., 2013).

Due to the Twomey effect, the shortwave cloud radiative effects are more

affected than longwave radiative effects (Alizadeh-Choobari and Gharaylou,

2017). With higher concentrations of aerosols, the number of condensation

nuclei or ice nucleation particles can rise, reducing precipitation and increasing

cloud cover and cloud lifetime (Albrecht, 1989). Global cloud albedo can

12

increase by means of the longer lifetime, higher top of clouds (Pincus and

Baker, 1994), and by smaller cloud droplets. Under a constant water content,

the increase of aerosol concentration can produce the decrease of cloud

droplets size and hence the increase of cloud albedo (Boucher et al., 2013).

This increase is more significant for low concentration of droplets (Andrejczuk et

al., 2014; Pincus and Baker, 1994). However, in some cases, with higher

number of larger aerosols (about 0.5 µm) the albedo can decrease (Andrejczuk

et al., 2014). Higher concentrations of aerosols increase the liquid water content

and the optical depth of cloud due to more condensation (Alizadeh-Choobari

and Gharaylou, 2017). High correlation between decreasing of precipitation and

load of aerosol dust has been reported around the world (Hui et al., 2008).

Alizadeh-Choobari and Gharaylou, (2017) reported lower cloud base under

pollution conditions due to the drier boundary layer as compared to clean

boundary layer.

Theoretically, absorption of solar radiation by aerosols induces a diabatic

heating and hence an increase of temperature, reducing the relative humidity

and delaying the saturation of the layer. The static stability of clouds can be

affected and the evaporation of cloud droplets occurs due to the heating,

decreasing cloud cover. This decrease allows for an extra warming of the

surface by solar radiation (Perlwitz and Miller, 2010). However it is not a simple

dynamic process, because when the layer is heating by absorption of radiation

by aerosols, the specific humidity increases. Thus the warmer layer produces a

moisture convergence. That can overcome the possible reduction of cloud

cover, increasing the cover of low clouds especially in summertime and over

land (Perlwitz and Miller, 2010). Those processes depend on a variety of factors

as atmospheric circulation and the influence of absorbing aerosols in the

specific humidity (Perlwitz and Miller, 2010).

2.3. Sao Paulo weather

Due to the geographical and topographic conditions, the weather in São

Paulo is strongly influenced by the synoptic systems and the local circulation of

breeze. In the austral summer, December to February, wet conditions prevail,

with local circulation and synoptic systems driving convective activity and hence

13

favoring cloud formation. The South Atlantic Convergence Zone (SACZ) affects

the region in summer with notable increase of cloud cover and flood events.

The SACZ can persist longer than 10 days, and usually more than 8 events

occur every year (Carvalho et al., 2004).

Since the Metropolitan Area of São Paulo (MASP) is located over a

plateau and about 50 km from the coast, the parallel orientation of the mountain

range, with mountain-valley circulation, contributes to the advance of sea

breeze inland. The sea breeze front can remain stationary over MASP by the

influence of heat urban island circulations (Freitas, 2003). The sea breeze front

passage occurs after midday, increasing the convection development in the

afternoon. This local condition is favored in summer due to the increase of

surface temperature and water vapor.

Silva Dias et al. (2013) discussed that, in the wet season, local factors

such as the increase of air pollution and the growth of Sao Paulo city might

have influenced the positive trend of rainfall in the last eight decades.

In winter, the region is affected by the passage of cold fronts and

migratory high pressure systems and, under this condition, imposing the

presence of cold dry air. Despite the fact that the highest frequency of cold

fronts occurs in August, in this season cold fronts are less rainy and move

faster, increasing cloud cover only near the coast (Morais et al., 2010). In

spring, the maximum number of cold fronts is observed with frequency around 8

days.

2.4. Shortwave radiative transfer

The sun emits energy in the form of electromagnetic waves, especially

those with wavelength below 4 µm. The sun electromagnetic spectrum is

divided in ultraviolet (UV) for wavelengths between 0.1 and 0.4 µm, visible (VIS)

above 0.4 µm up to around 0.7 um and near infrared (NIR) between 0.78 µm

and 3.5 µm (Yamasoe and Corrêa, 2016).

Radiation can be quantified by energy per unit of time, the so called

radiant power. Considering a unit of surface area, a pencil of radiation beam

(given by infinitesimal solid angle) with an orientation defined by the cosine of

14

the zenith angle (𝜇) (relative to the vertical axis) also referred as CZA, and

azimuth angle (𝜙) (angle along the horizontal axis) is named as radiance. Thus,

radiance is radiant power per unit area orthogonal to the pencil of radiation, per

unit of solid angle. The integration of all pencils of radiation in the given solid

angle (in general, in one hemisphere) is the irradiance, considering a horizontal

surface. Radiance is expressed in Wm-2sr-1 and irradiance in Wm-2.

Through the atmosphere, solar radiation undergoes extinction processes by

means of absorption and scattering. If the solar beam suffers any extinction due

to the scattering, the scattered component is named diffuse irradiance or

radiance; otherwise it is called as direct sun radiance or irradiance.

Scattering is a process of deviation of radiance due to interaction of the

electromagnetic field of radiation and the electromagnetic field generated by

particles. The scattering does not involve the transformation of energy and it is

sensitive to the relationship between the size of the scattering particle and the

wavelength of radiation. For particle much smaller than the wavelength of

radiation, the scattering is symmetric and spectrally sensitive (Rayleigh

scattering). When the particle size is comparable to the wavelength of the

incident radiation, anisotropy in the scattered field is observed, theoretically

explained by Mie Theory (Liou, 2002). In the Mie scattering, there is less

spectral dependence and the forward scattering is more pronounced.

Molecular absorption is a consequence of the interaction of the electrical

dipole of the molecule with electromagnetic radiation. Therefore, it depends on

the geometric distribution of electrical charges of the molecule and the

wavelength where vibrational and or rotational modes of the molecule are

activated. Because of that, absorption of gases is highly wavelength dependent

(Liou, 2002). For particles, the absorption depends on the imaginary part of the

refractive index and on the relationship of the particle size to the wavelength of

the incident radiation and, opposite to gases, it is spectrally continuous (Liou,

2002).

All these processes can be quantified by the shortwave radiative transfer

equation (RTE). Considering a plane parallel atmosphere for spectral diffuse

15

radiance 𝐼𝜆 at any layer with height (z) and orientation (±µ, ϕ), the RTE is shown

in equation 1 (Yamasoe and Corrêa, 2016), where cos Θ0 represents the solar

beam direction and cos Θ′ represents all possible incident directions of diffuse

radiation. Note the dependence on single scattering albedo (𝜔, also known as

SSA), phase function (𝑝) and extinction coefficient (𝛽𝑒) which are the optical

properties characterizing the atmospheric layer. The second term on the right

represents diffuse radiance generated by single scattering process (only one

scattering of the solar direct beam), while the third term indicates scattering of

diffuse radiance already available at that atmospheric layer (multiple scattering)

to the observer direction (±µ, ϕ).

±𝜇𝑑𝐼𝜆(𝑧,±𝜇,𝜙)

𝛽𝑒𝜆(𝑧)𝑑𝑧= 𝐼𝜆(𝑧, ±𝜇, 𝜙) −

𝜔𝜆(𝑧)

4𝜋𝑆𝜆(𝑧)𝑝𝜆(𝑧, cos Θ0) −

−𝜔𝜆(𝑧)

4𝜋∫ ∫ 𝐼𝜆(𝑧, ±𝜇′, 𝜙′)

1

−1

2𝜋

0𝑝𝜆(𝑧, cos Θ′) 𝑑𝑢′ 𝑑𝜙′ (1)

The extinction coefficient is the sum of absorption (𝛽𝑎) and scattering (𝛽𝑠)

coefficients and can be computed from a volume containing particles or

molecules (equation 2), where 𝑛(𝑧) is the density number in any determined z

and 𝜎𝑒,𝑠,𝑎(𝑧) denotes the cross section due to extinction, scattering or

absorption. The 𝜎𝑒,𝑠,𝑎 (𝑧) is calculated from the extinction, scattering or

absorption efficiency (𝑄𝑒,𝑠,𝑎(z)) and the geometrical cross-section (A).

𝑄𝑒,𝑠,𝑎 quantifies the extinction, scattering or absorption per cross sectional area

unit, and is calculated from Mie Theory for aerosols and clouds. For gases,

Rayleigh scattering theory is used to compute efficiencies (Liou, 2002). For

absorption due to gases, equation 4 is used, where 𝑘 is the mass absorption

coefficient and 𝜌𝑎 the density of the gas.

𝛽𝑒,𝑠,𝑎𝜆 (𝑧) = 𝜎𝑒,𝑠,𝑎𝜆 (𝑧)𝑛(𝑧) (2)

𝜎𝑒,𝑠,𝑎𝜆(𝑧) = 𝑄𝑒,𝑠,𝑎𝜆

(𝑧) 𝐴 (3)

𝛽𝑎𝜆 (𝑧) = 𝑘𝜆𝜌𝑎(𝑧) (4)

𝜏𝜆 = ∫ 𝛽𝑒𝜆(𝑧)𝑑𝑧

𝑧′

∞ (5)

16

𝜔𝜆(𝑧) =𝛽𝑠𝜆

(𝑧)

𝛽𝑒𝜆 (𝑧) (6)

𝑝𝜆(cos Θ) = ∑ (2𝑖 + 1) 𝑋𝑖 (𝑃𝑖(cos Θ))2𝑁−1𝑖=0 (7)

𝑝𝐻𝐺𝜆(cos Θ , 𝑔𝜆) =

1−𝑔𝜆2

(1+𝑔𝜆2−2𝑔𝜆 cos Θ)

32

(7.b)

𝑔𝜆 =1

2∫ 𝑝(cos Θ)

1

−1 cos Θ 𝑑 (cos Θ) (8)

When 𝛽𝑒𝜆is integrated along z from TOA up to the z’ level, the optical

depth of the medium is obtained (equation 5). This equation can be used for

computing the optical depth of each atmospheric component if the 𝛽𝑒𝜆due to

this component is known.

The single scattering albedo is a ratio that represents the fractional part

of the extinction process related to the scattering (equation 6), i. e. if 𝜔 is 1 the

extinction is only due to scattering, otherwise if 𝜔 is 0 the extinction is only a

consequence of absorption. The phase function (𝑝) gives the probability of

redistribution of scattered radiation. It can be computed with high accuracy

using Legendre polynomials (𝑃𝑖), according to equation 7 (Yamasoe and

Corrêa, 2016). Legendre coefficients of each expansion are 𝑋𝑖 and the number

of terms N. The higher the size of particle the more complex the computation of

𝑝. Due to the computational cost, phase function can be computed by analytic

functions such as the Henyey-Greenstein approximation, given in the equation

7b. The main input variable of Henyey-Greenstein approximation is the

asymmetry parameter (g), the first coefficient 𝑋0 (equation 8). It represents the

degree of anisotropy of the phase function e .g. for Rayleigh scattering g=0 and

for total forward scattering g=1 (Yamasoe and Corrêa, 2016).

The direct component of RTE (equation 9) quantifies the radiative

transfer of the solar beam through a medium characterized by an extinction

coefficient βeλ. It is also known as the Beer-Lambert-Bouguer law (equation 10).

The 𝑆0𝜆 is the radiance at TOA at the standard distance of 1 unit astronomical

distance between the Earth and the Sun, 𝑈 is the correction factor of Earth-Sun

distance and 𝜇0 the cosine of zenith angle in the direction of the solar beam.

Equation 10 is widely used for retrieving optical properties of aerosols, water

17

vapor and even thin clouds using direct sun measurements (Liou, 2002; Min

2004)

The total optical depth is the sum of optical depths of every component.

In equation 11, 𝜏𝑟𝑎𝑦 is the optical depth by Rayleigh scattering of gas

molecules. Optical depth due to absorption by gases (𝜏𝑎𝑏𝑠) is the sum of optical

depths of absorbing gases such as ozone, water vapor, carbon dioxide etc. In

the case of semitransparent clouds blocking the solar beam, the optical depth of

clouds is 𝜏𝑐𝑙𝑑 or COD. Usually, in the presence of thick clouds, the solar direct

beam is completely attenuated, with only the diffuse component reaching the

surface.

The monochromatic transmittance (𝑇𝑑𝑖𝑟𝜆) is defined by the ratio between

direct radiance at any z level 𝑆𝜆 (𝑧) and 𝑆0𝜆 (equation 12). The term 𝜇0−1 is also

known as the relative optical air mass of the layer (𝑚), but only in the plane-

parallel approximation.

−𝜇0𝑑𝑆𝜆(𝑧)

𝛽𝑒𝜆(𝑧)𝑑𝑧= 𝑆𝜆 (9)

𝑆𝜆(𝑧) = 𝑆0𝜆𝑈 exp−𝜏

𝜇0= 𝑆0𝜆𝑈 exp( −𝑚𝜏) (10)

𝜏𝜆 = 𝜏𝑟𝑎𝑦𝜆+ 𝜏𝑎𝑏𝑠𝜆

+ 𝜏𝑎𝑒𝑟𝜆+ 𝜏𝑐𝑙𝑑𝜆

(11)

𝑇𝑑𝑖𝑟𝜆(𝑧) =

𝑆𝜆(𝑧)

𝑆0𝜆𝑈= exp

−𝜏𝜆

𝜇0 (12)

Using equations 13 and 14, the monochromatic downwelling and

upwelling diffuse irradiance are computed, respectively. The spectral direct

solar irradiance is 𝑆𝑆𝜆 is computed integrating all the radiances in the solid

angle of the solar disk (Ω𝑠𝑢𝑛) (equation 15). The global irradiance (𝐺) at any

height is the sum of downwelling diffuse and direct sun irradiance orthogonal to

the surface (equation 16). Thus, the hemispherical reflectance (R) is defined as

the ratio of upwelling irradiance and 𝐺 at the same height level (equation 17).

The total transmittance 𝑇at any level is calculated from equation 18 as the ratio

of G of the level and the global irradiance at TOA (𝐺(∞)). Note, 𝐺(∞) is

18

computed from direct sun irradiance at TOA in 1 astronomical unit (𝑆𝑆𝜆0)

corrected by the distance factor (𝑈) and by 𝜇0 or CSZA (equation 19).

𝐷 ↓𝜆 (𝑧) = ∫ ∫ 𝐼 (𝑧, −𝜇, 𝜙)𝜇𝑑𝜇0

1

2𝜋

0𝑑𝜙 (13)

𝐷 ↑𝜆 (𝑧) = ∫ ∫ 𝐼 (𝑧, 𝜇, 𝜙)𝜇𝑑𝜇0

−1

2𝜋

0𝑑𝜙 (14)

𝑆𝑆𝜆(𝑧) = ∫ 𝑆𝜆(𝑧, 𝜇, 𝜙) dΩ(𝜇, 𝜙)Ω𝑠𝑢𝑛

(15)

𝐺𝜆(𝑧) = 𝑆𝑆𝜆(𝑧)𝜇0 + 𝐷 ↓𝜆 (𝑧) (16)

𝑅𝜆(𝑧) =𝐷↑ 𝜆(𝑧)

𝐺𝜆(𝑧) (17)

𝑇𝜆(𝑧) =𝐺𝜆(𝑧)

𝐺𝜆(∞) (18)

𝐺𝜆(∞) = 𝑆𝑆𝜆0𝑈𝜇0 (19)

Hereinafter the downwelling broadband diffuse irradiance is defined as D

and direct sun irradiance as SS when they are integrated over λ. For specific

wavelength a subscript will be employed.

2.4.1. Cloud optical and radiative properties

Effective radius (re) and liquid/ice water content (LWC/IWC) are the cloud

properties better related to the cloud optical properties (𝜔, 𝑔 and COD). The re

for liquid clouds is computed as the ratio of the integrated third and second

moments of size distributions (equation 20). It represents the mean radius

weighted by the cross sectional area of drops. LWC is the sum of the mass of

all drops in the distribution (equation 21). For this parameter, the knowledge of

the number distribution 𝑁(𝑟) and the density of liquid water (𝜌) is necessary

(Yamasoe and Corrêa, 2016).

In the shortwave spectrum, cloud optical properties are insensitive to the

shape of cloud drop size distribution and mainly depend on re, LWC or IWC and

wavelength (Hu and Stamnes, 1993). The 𝜔 and g only depend on re and

wavelength (Rawlins and Foot, 1989). Spectral optical properties of clouds in

19

the homogenous plane parallel conditions are frequently characterized by re and

COD.

re =∫ 𝜋𝑟3𝑁(𝑟) 𝑑𝑟

∞0

∫ 𝜋𝑟2𝑁(𝑟) 𝑑𝑟∞

0

(20)

𝐿𝑊𝐶 =4𝜋

3𝜌 ∫ 𝑟3𝑁(𝑟) 𝑑𝑟

0 (21)

Because the size of cloud drops is larger than the wavelength of solar

radiation, The Mie Theory is considered for computing cloud optical properties

taking into account the refractive index and radius of particle. For liquid clouds

𝛽𝑒,𝑠,𝑎 is calculated by integrating the 𝑄𝑒,𝑠,𝑎 over drop size distribution (𝑛(𝑟))

(equation 22).The 𝑄 depends on wavelength (𝜆), refractive index (ni) and radius

(r) of given cloud particle.

𝛽𝑒,𝑠,𝑎𝜆= ∫ 𝜋𝑟2𝑄𝑒,𝑠,𝑎𝜆

(𝑟, 𝑛𝑖)𝑛(𝑟)𝑑𝑟∞

0 (22)

Since 𝑄𝑒 tends to a constant value of 2, for larger size of cloud drops as

compared to the wavelength of solar radiation, it is possible to derive a simple

relationship between 𝛽𝑒, re and LWC (equation 23):

𝛽𝑒 ≈3 𝐿𝑊𝐶

2𝑟𝑒 (23)

Scattering and absorption properties of ice particles cannot be computed

using the Mie Theory because of the non-spherical characteristics of ice

particles, instead ray tracing technique is employed (Yang et al., 2000). For ice

clouds re is obtained using the maximum dimension (L) of ice particle (equation

24), where V and A are the equivalent volume and the projected area assuming

a spherical particle (Yang et al., 2000). The IWC can be retrieved from the

distribution number of particles in function of L, 𝑁(𝐿), by equation 25, assuming

the density of ice (𝜌𝑖𝑐𝑒) as 0.9167 g.cm-3 (Key et al., 2002).

𝑟𝑒𝑖𝑐𝑒 =

∫ 𝑉(𝐿)𝑁(𝐿)𝑑𝐿∞

0

∫ 𝐴(𝐿)𝑁(𝐿)𝑑𝐿∞

0

(24)

𝐼𝑊𝐶 = 𝜌𝑖𝑐𝑒 ∫ 𝑉(𝐿)𝑁(𝐿)𝑑𝐿∞

0 (25)

20

In the Figure 2 the spectral optical properties estimated using the

parameterization developed for liquid clouds (Hu and Stamnes, 1993) and for

rough-aggregated "habit" of ice clouds of Key et al. (2002) are shown. For liquid

clouds, optical properties are presented for re values of 5 µm, 10 µm and 15 µm.

In the case of ice clouds they are computed for re equal to 30 µm, 51.2 µm and

70 µm. Differences of g between clouds are observed, with higher values for

liquid clouds in the spectral range from 0.2 µm up to 1.5 µm (Figure 2.a). Thus,

for the same cloud optical depth, for liquid clouds more forward scattering and

hence a higher transmission of solar radiation is expected as compared to ice

rough-aggregated (near 3D ice) (LeBlanc et al., 2015). However, for plates and

columnar ice particles, g can be higher than for liquid clouds (Key et al., 2002).

Figure 2: Asymmetry parameter (g) (a), single scattering albedo (𝝎) and the ratio between the

extinction coefficient and LWC or IWC (𝛽𝑒 /LWC or IWC) for liquid and ice clouds. For liquid clouds, Hu and Stamnes (1993) parameterization is employed and for ice clouds, the parameterization of Key et al. (2002) for rough-aggregate habit is used. The optical properties are computed for re values of 5,10 and 15 µm for liquid clouds and for 30, 51.2 and 70 µm for ice clouds.

Asymmetry parameter g is sensitive to re (Figure 2.a), while ω near 1

confirms that for clouds, scattering dominates in wavelengths below 1.1 µm.

The absorption of radiation by clouds (ω below 1) is observed for wavelength

higher than 1.1 µm (Figure 2.b) (Marshak et al., 2000). In the solar spectrum,

the extinction coefficient is almost spectrally constant, especially for larger re

(Figure 2.c). 𝛽𝑒 decreases with re and strongly depends on LWC or IWC.

Although 𝑄𝑒 converges to 2 for large size parameter (2𝜋𝐿𝜆⁄ ) also for ice

clouds (Yang et al., 2000), the optical properties of ice clouds depend on the

particle habit. Differences are significant for g in the visible spectrum. Larger ice

21

plates have strong forward scattering with values of g around 0.94. Solid how

columns and bullet rosettes have maximum values of g near 0.86 while for

aggregate, the maximum hardly reaches 0.8. Thus, in the VIS, the R of cloud is

higher (0.65 for COD of 6) for aggregates and solid columns and lower (0.52 for

COD of 6) for ice plates (Key et al., 2002).

The R of clouds is more sensitive to re than T, especially in spectral

regions where the single scattering albedo of clouds is less than 1 (Rawlins and

Foot, 1989). That is because variations of re lead to g and 𝜔 or SSA responses

in the same direction. i. e., the increase (decrease) of re increases (decreases)

the absorption efficiency and forward scattering, contributing to the decrease

(increase) of R. Because of that, retrieving re by measurements of solar

radiation reflected by clouds onboard satellites is more effective (McBride et al.,

2011). At the surface, the opposite behavior of SSA and g causes the less

sensitivity of T to re. In the NIR, larger re results in lower SSA values, reducing

T. The increase of the forward scattering, for larger re, results in the increase of

T. Thus, T measured at the surface in the presence of clouds is more sensitive

to COD than to re (Min and Harrison, 1996). For wavelengths below 1.1 µm, the

higher the re leads to the increase of radiance at zenith due to the direct impact

of the forward scattering. By contrast, for wavelength about 1.4 µm, as the

absorption process becomes more important, the transmitted radiance at

zenith decreases (McBride et al., 2011).

The first works studying cloud and solar radiation interactions focused on

the bulk radiative properties of clouds, i.e reflectivity known as cloud albedo,

absorptivity (the amount of solar radiation absorbed by the cloud layer) and

transmissivity.

One of the first results in the scientific literature about the cloud effects

on solar radiation appeared in the 40’s of the last century (Neiburger, 1949).The

author computed bulk radiative properties of clouds using downwelling and

upwelling shortwave radiation measurements below, inside and above the

coastal stratus clouds in the United States. He showed that the cloud effects

depended on cloud thickness and confirmed, by measurements that the cloud

albedo is the most important property. In addition, low cloud absorption was

22

observed. In that epoch, they argued the need of increasing observations of

cloud microphysical properties.

Paltridge (1974) combined radiative and microphysical measurements

made in stratocumulus clouds. He observed the increase of LWC at high layers

of clouds near the cloud top, with the strong albedo relationship with LWC,

therefore the highest albedo was observed near the cloud top.

Manton (1980) showed the increase of cloud albedo with drops number,

which is higher for thin clouds. Differences were also observed between

maritime and continental clouds, with the continental reflecting 5 % more than

maritime clouds. He also showed the decrease (increase) of reflectivity with re

(LWC). The higher reflectivity near cloud top was strongly related with the

increase of density in this part of the cloud. In addition, the reflectivity

(transmissivity) increased (decreased) for lower CSZA especially lower than

0.5.

Ackerman and Stephens (1987) argued that the scattering properties of

clouds do not depend on droplet size distribution. The cloud absorption

efficiency decreases as an inverse function of re for thin clouds. For deeper

clouds they showed an opposite behavior, with increasing absorption efficiency

for higher re.

2.4.2. Retrieving COD and re from ground based

measurements

Ground-based measurements are useful sources of data to study the

radiative processes in the atmosphere and for validation of satellite

measurements. Around the world, there are still few sites, especially in the

southern hemisphere. A variety of methods for computing cloud optical

properties from passive ground-based measurements has been developed in

the last two decades.

Two approaches for retrieving COD from passive ground-based

instruments are reported in the literature: using broadband or narrowband G,

23

the so called effective cloud optical depth (ECOD) and COD retrieved locally by

using zenith radiance measurements. The former works well for clouds with

total overcast conditions and using 1D radiative transfer model, because the

one to one relationship can be observed between irradiance and COD. In the

case of broken cloud fields, each element of the sky contributes differently to

the irradiance (Marshak et al., 2004), even with enhancement effects, due to 3D

effects of clouds.

Retrieving COD locally has the disadvantage of no one to one

relationship between zenith radiance and COD. Therefore, using combinations

of spectral radiance measurements is needed (Brückner et al., 2014; Chiu et al.,

2010; Marshak et al., 2004). In the following, examples of methods found in the

literature are described.

Leontyeva and Stamnes (1994) proposed the COD estimation from

transmitted irradiance using a broadband pyranometer. They projected a simple

model assuming plane parallel homogenous clouds, without considering the

cloud type or cloud base height. The possible influence of water vapor is not

well represented. They suggested the use of this methodology, but employing a

more accurate model and measurements in the UV and VIS to avoid the

influence of water vapor absorption in the infrared.

Barnard and Long (2004) proposed a simple method for the estimation

of COD using a relationship of COD to G with cases of clouds with total

overcast conditions. They did not consider in depth the atmospheric conditions

for the retrieval; therefore it is less accurate than methods employing spectral

measurements. They also suggested avoiding the use of the method when high

accuracy is needed.

Min and Harrison, (1996) retrieved COD of warm clouds in total overcast

conditions using T415 and surface albedo computed from a Multi-Filter Rotating

Shadowband Radiometer (MFRSR). They also retrieved re from vertical liquid

water path measured by zenith-viewing microwave radiometer (MWR). The

retrieval employed a 1-D radiative transfer model based on the discrete ordinate

method. The advantage of using measurements at channel centered at 415 nm

instead of other spectral regions is argued: the lower surface albedo, the less

24

sensitivity of SSA and g to re and because the absorption of radiation by ozone

in the Chappuis-band can be avoided. COD is retrieved by nonlinear square

method considering cloud properties as stationary in fixed time intervals, by

minimizing the sum of errors in transmittance. The poor ability to distinguish

higher optical depths from satellite measurements at the top of the atmosphere

and the lower uncertainties of retrieving COD from ground-based

measurements were demonstrated.

The best accuracy, close to 5 %, of COD retrieved for thin clouds from

MFRSR was obtained by Min (2004). The method used direct sun irradiance

obtained from MFRSR measurements. A correction for the forward scattering to

avoid underestimation due to the increase of the diffuse contribution in the

direct orientation was developed. The method employed the Bouguer-Lambert-

Beer Law (see equations 10-11) and aerosols can be discriminated from clouds,

using the spectral relationship between the optical depth at 415 nm and 860

nm, with the Ångström exponent (equation 37). It is only applied to thin clouds,

when the direct sun irradiance can be measured.

On the other hand, Marshak et al. (2000) proposed a method for locally

retrieving COD of low broken clouds above green vegetation, using the spectral

contrast of albedo in VIS and NIR for vegetated surface in the presence of low

liquid clouds. COD of clouds is almost constant in the shortwave spectral range.

Therefore in the presence of clouds, radiance at zenith is strongly influenced by

multiple scattering between surface and cloud base. They used the spectral

radiance at 650 nm and 870 nm normalized to the radiance at the top of the

atmosphere. By analogy with the normalized difference vegetation index

(NDVI), they defined the Normalized Difference Cloud Index (NDCI). The NDCI

index minimizes the enhancement effect of zenith radiance due to broken low

cloud fields and shows a good agreement with COD for horizontal resolution

higher than 0.4 km. In addition, the method does not take into account the cloud

fraction, that can lead to poor retrievals.

Based on the relationship of COD with spectral radiance previously

explained, Chiu et al. (2010) proposed a method to compute COD applied to

AERONET measurements, in conditions when aerosol properties cannot be

25

retrieved, called the cloud mode. They used zenith radiance between 440 nm

and 870 nm to take the advantage of more spectral contrast of surface albedo.

This method is effective for low broken cloud fields and it is very sensitive to

aerosol properties, but the authors did not show the errors due aerosols,

arguing that COD retrievals frequently are higher than 15. The total error of the

retrieval is about 17 %, with 4 % due to the re error of 25 %.

McBride et al. (2011) developed a method to retrieve COD and re from

zenith transmitted radiance at 515 nm and the slope of transmitted radiance at

1565 nm and 1634 nm normalized to radiance transmitted at 1565 nm. The

method takes advantage of the variability of spectral absorption efficiency of

clouds in the NIR to obtain the best accuracy for retrieving re. That resulted in

higher sensitivity of the normalized transmittance at NIR to re.

Differences in the shortwave spectral absorption due to the cloud phase was

taken into account by Le Blanc et al. (2015). They derived 15 parameters to

obtain COD, re and cloud phase from spectral transmittance given the cloud

type. But only 2 parameters are derived in the spectral range of VIS, with the

rest in the NIR.

Brückner et al. (2014) retrieved COD and re from 3 ratios of transmissivity

using six spectral ranges (450nm, 680 nm, 1050nm, 1250 nm, 1670 nm, 1560

nm) of zenith radiances. The method is applied to homogenous and

heterogeneous liquid water clouds and cirrus clouds using a 1-D radiative

transfer model. Using combinations of these transmissivity ratios, problems

related to the retrievals of COD lower than 5 were reduced. Errors of retrievals

were near 5 % for low clouds and 10 % for ice clouds, increasing for thicker

clouds.

2.4.3. Computing the shortwave radiative effects of clouds

The shortwave cloud radiative effect (CRE) is a variable that quantifies

the amount of net solar irradiance, or G, affected by the influence of clouds at a

horizontal surface. It can be computed using the 𝐷 ↑𝑐𝑙𝑑 and G (Salgueiro et al.,

2016) in equation 26 or only using G (Tzoumanikas et al., 2016) in equation 26.

In the equations 26 and 28, the subscript clear represents computations under

26

clear sky conditions; usually a radiative transfer model is employed to obtain

this value. The subscript cld corresponds to the irradiances measured in cloudy

conditions. Sometimes 𝐷 ↑𝑐𝑙𝑑 is not measured directly, therefore the use of

surface albedo (𝛼), is needed as shown in equation 27. By the equation 28,

CRE can be computed not including the surface albedo. The CRE is given in W

m-2 and negative values indicate a cooling effect, due to extinction of irradiance

by clouds. Positive values confirm the enhancement effect or the increase of

solar radiation at the surface due to 3D scattering effects by clouds.

𝐶𝑅𝐸 = 𝐺𝑐𝑙𝑑 − 𝐷 ↑𝑐𝑙𝑑− 𝐺𝑐𝑙𝑒𝑎𝑟 + 𝐷 ↑𝑐𝑙𝑒𝑎𝑟 (26)

𝐷 ↑𝑐𝑙𝑑= 𝛼 𝐺𝑐𝑙𝑑 (27)

𝐶𝑅𝐸𝐺 = 𝐺𝑐𝑙𝑑 − 𝐺𝑐𝑙𝑒𝑎𝑟 (28)

The CRE decreases with CSZA and increases with cloud cover

(Tzoumanikas et al., 2016). Other variables are used instead of CRE in order to

minimize the CSZA effect: the normalized cloud radiative effect (NCRE)

(Salgueiro et al., 2016) (equation 29) and cloud modification factor (CMF)

(equation 30). They are variables computed as the ratio of G in all sky

conditions to G in clear sky ones, and they almost totally eliminate the solar

zenith angle effect.

𝑁𝐶𝑅𝐸 =𝐶𝑅𝐸𝐺

𝐺𝑐𝑙𝑒𝑎𝑟 (29)

CMF =𝐺𝑐𝑙𝑑

𝐺𝑐𝑙𝑒𝑎𝑟 (30)

2.4.4. Cloud radiative efficiency (CEF)

To quantify how the cloud radiative effect changes per COD unit, the

cloud radiative efficiency (CEF) is defined. The CEF is computed from the

relationship between the cloud radiative effect and COD. In the case of cloud

with COD below 3 a linear relationship between CRE and COD for any CSZA

are observed, but for higher values of COD an exponential relationship is

expected and CEF is computed from linear relationship between CRE and

ln(COD) for ranges of CSZA (Mateos et al., 2014).

27

In equations 31-33, CEF can be estimated according to Mateos et al.,

(2014) where b is the slope of the relation for different values of CSZA. CEF

computed from CRE has unit of Wm-2 (COD-unit)-1 (Mateos et al., 2014), while

NCRE is in (COD-unit)-1 or % (COD-unit)-1 if NCRE is given in percent values

(Salgueiro et al., 2016).

𝐶𝑅𝐸 = 𝑎 + b ln( 𝐶𝑂𝐷) (31)

𝑏 =Δ𝐶𝑅𝐸

Δ𝐶𝑂𝐷𝐶𝑂𝐷 (32)

𝐶𝐸𝐹 =Δ𝐶𝑅𝐸

Δ𝐶𝑂𝐷=

𝑏

𝐶𝑂𝐷 (33)

CEF is sensitive to CSZA and COD when it is computed from CRE

(Mateos et al., 2014), it increases with CSZA and decreases with COD. Mateos

et al. (2014) showed a maximum efficiency of -160 W/m2 per COD unit at high

CSZA and low COD and minimum of -0.3 W/m2 per COD unit when CSZA is

lower with moderate-large COD. They retrieved COD using broadband

pyranometer measurements with less accuracy, showing a strong relationship

between COD and CRE for a fixed CSZA. In contrast, (Salgueiro et al., 2016)

achieved the best accuracy when CEF was computed based on the linear

relationship between NCRE and ln(COD).

2.4.5. The enhancement effect of clouds on the solar

radiation

In the presence of clouds, if the solar disk is not obstructed, the direct

solar irradiance is not attenuated and the diffuse irradiance can increase. Due

to this effect, called enhancement, the solar G measured at the surface can be

higher than the solar G measured under clear sky conditions keeping all other

variables identical.

The increment of diffuse irradiance is whether can be a consequence of

the multiple reflection of the direct solar irradiance in the border of clouds or the

increase of forward scattering by the cloud (Tzoumanikas et al., 2016). In

special cases, G can overcome an equivalent G at the top of the atmosphere.

28

According to Almeida et al. (2014) these events are called as overirradiance

events.

Tzoumanikas et al (2016) argued that the necessary conditions for cloud

solar enhancement events are: the presence of cumulus clouds, very dense but

not blocking the solar disk, with a minimum value of 50 % and maximum of 90

% cloud cover. These events can last from few seconds to minutes and, in

extreme cases, the measured downward irradiance can be higher than 1500

W/m2 (Tzoumanikas et al., 2016). They observed maximum enhancement of

200 W/m2, occurring in the presence of high and mid-level clouds, values

comparable to low clouds. The position of the solar disk relative to the clouds is

a crucial condition for strong enhancement: solar disk near clouds or low CSZA

with clouds present near zenith (Tzoumanikas et al., 2016). Marín et al. (2017)

showed that the enhancement can be observed at any time of the day CSZA

and the maximum is observed between 40-50 % of cloud cover. They reported

a median of NCRE of 0.18 and percentile 75 of 0.3 at Valencia, Spain.

Enhancement effects at São Paulo were reported before, with the presence of

overirradiance events with time resolution shorter than 1 minute, which were

related to broken cumulus (Almeida et al., 2014).

2.4.6. The shortwave attenuation effect of clouds

The shortwave cooling effect of clouds at the surface is determined when

values of CRE or NCRE are negative. It represents the reduction of solar

radiation at the surface by clouds, as compared to clear sky at the same time.

Tzoumanikas et al (2016) observed maximum of CRE near -900 W/m2

computed with surface albedo. They reported in Greece, strong cooling effects

due to Sc and Cb near -800 W/m2 with CSZA around 0.86. Cu has cooling

effects as high as Sc, but a high variability of CRE was reported with many

cases below -200 W/m2. In the tropical Western Pacific region, Burleyson et al.

(2015) reported the strongest cooling for Cb with NCRE values between -0.81

and -0.84, for low clouds between -0.37 and -0.48, cirrus and mid clouds

presented mean values of NCRE comparable to boundary layer low clouds with

mean between -0.17 and -0.57 and -0.32 and -0.44, respectively.

29

At São Paulo, Yamasoe et al. (2017) reported higher cooling effects

computed using 24 h mean irradiance in summer of -170 Wm-2 and in winter of -

50 Wm-2. They computed mean values of CMF of 0.3 and 0.8 for low and high

clouds respectively.

30

3. Instruments and methods

3.1. Satellite data

3.1.1. Description of satellites

The A-train satellite afternoon constellation is a group of polar orbiting

satellites sharing the same orbit one after another. The time lapse between the

first and the last satellite is only around 15 minutes. Inside this constellation

CALIPSO and CloudSat, launched in 2006, make measurements with intervals

of 15 seconds. The constellation is known as the afternoon due to its overpass

over the equator at 13:30 local time. Near São Paulo, the overpass occurs in

the afternoon around 14:00 LT and near 01:00 LT at night.

CloudSat (acronym of Cloud Satellite) allows estimating the vertical

structure of clouds and precipitation. The main instrument onboard is the

microwave radar with nadir view called Cloud Profiling Radar (CPR) that

measures backscattering from cloud particles at 94 GHz (3.2 mm). The footprint

at sea level has a resolution of 1.7 km and 1.3 km along and across track,

respectively. The vertical resolution of bins is about 240 m and the minimum

detectable radar reflectivity factor is approximately −30 dBZ. Because of the

wavelength of 3.2 mm, CPR of CloudSat is less sensitive to smaller cloud drops

(subvisual cirrus or shallow water droplets), but it can penetrate optically thick

clouds. Moreover, no valid measurements near the surface, specially under 1

km, can be made, due to the backscattering contamination from the surface

(Marchand et al., 2008). The CPR only detects 30 % of clouds under 1 km and

fails in 40 % for retrieving cloud properties due to the limitations of CPR

(Christensen et al., 2013). Since April of 2011, measurements by the satellite

are made only during daytime, because of a battery anomaly.

CALIPSO (acronym of the Cloud-Aerosol Lidar and Infrared Pathfinder

Satellite Observations) satellite has aboard the Cloud-Aerosol Lidar with

Orthogonal Polarization (CALIOP). From CALIOP measurements,

backscattering coefficients at 532 nm and 1064 nm and depolarization ratio at

532 nm are retrieved. The main vertical and horizontal resolutions are 30 m and

333 m, respectively up to 8.2 km, and from 8.2 km until 20 km the vertical and

31

horizontal resolutions are 60 m and 1 km respectively (Winker et al., 2009).

With CALIPSO, smaller cloud drop such as from sub visual cirrus or

semitransparent clouds of boundary layer, can be detected. Hence, the synergy

of radar and lidar is very powerful because the constraints of each instrument

on detecting clouds are minimized.

3.1.2. CloudSat and CALIPSO products

In this research, microphysical variables for liquid clouds are obtained

from 2B-CWC-RO products of CloudSat. The data can be downloaded in the

following website: http://www.cloudsat.cira.colostate.edu/). In addition, because

of this product also has profiles for liquid clouds available, cloud properties such

as cloud thickness can be estimated for low clouds.

For ice clouds, properties are obtained from CSCA-Micro of JAXA

products (http://www.eorc.jaxa.jp/EARTHCARE/research_product/ecare_monito.html).

Cloud particle definition from CA-Ctype was used to estimate the frequency of

habit for layer and for clouds. Cloud mask of lidar/radar from CSCA-Cmask was

employed to compute the geometric properties cloud base for liquid and ice

clouds and thickness for ice clouds. Cloud frequency by height level was also

obtained for each season and according to distance from the coastal line. The

coastal line distance was estimated as the minimum distance of the

measurement point to the coast. The computations of frequency of occurrence

are valid if at least 100 measurements were found in every layer, or every bin

selected of coastal distance.

The time period of 2B-CWC-RO spans from December 2007 to June

2016 with a total of 436 overpasses. Quality control was made neglecting

measurements under precipitation and uncertainties above 20 %. In addition, a

total of 382 overpasses were found in JAXA products in the full period between

July 2006 and August 2015. In the following, the methodology of retrieving each

product is briefly described.

3.1.2.1. CloudSat

32

Liquid water content (LWC), ice water content (IWC), effective radius (re)

are included in the Cloud Water Content using Radar Only product (2B-CWC-

RO) from CloudSat products. The retrieval algorithm uses at first a cloud mask

from 2B-GEOPROF product to detect the layer with presence of clouds (Mace,

2007). Later, it employs 2B-CLD-CLASS product where cloud types are defined

and allows for checking if a level previously classified as cloudy is undetermined

or invalid which indicates troubles in the vertical profile of clouds (Austin, 2007).

The algorithm assumes the entire profile with ice or water clouds.

Finally, microphysics properties are retrieved using a forward scattering

model using a priori information of size distribution of drops. The model

considers lognormal size distribution depending on three parameters. In

addition, the radar reflectivity, LWC/IWC and re are related with the same three

parameters. Thus, the final retrieval is made by iteration from a priori

information up to the convergence of the solution (Austin, 2007). The algorithm

underestimates re and LWC/IWC in the presence of precipitation, dBZ>-15,

because it does not consider the size distribution of precipitating drops (Austin,

2007). Christensen et al. (2013) showed the high uncertainty in the retrieval for

mixed-phase profile due to the erroneous selection of refractive index. 2B-

CWC-RO has a horizontal resolution of 1.1 km with 125 vertical levels of 240 m

each.

3.1.2.2. CloudSat-CALIPSO (JAXA products)

Recently, a synergy between CALIPSO-CloudSat was publicized by the

EarthCARE Research Product Monitor of Japan Aerospace Exploration Agency

(JAXA), where between others, microphysics properties of ice clouds, a new

cloud mask and cloud type particle or habit of ice clouds are available. Products

of JAXA Radar/Lidar Cloud Mask Product (CSCA-Cmask), Lidar Cloud Particle

Type Product (CA-Ctype) and Radar/Lidar and Cloud Microphysics Property

Product (CSCA-Micro) were employed in this research.

CSCA-Cmask includes a new cloud mask from merged CALIPSO-

CloudSat cloud masks. This new cloud mask developed by Hagihara et al.

(2010) incorporated and improved cloud mask from CALIPSO and cloud mask

from 2B-GEOPROF of CloudSat. They applied two criteria for defining a cloudy

33

bin: the first considering a threshold of total attenuated backscattering

coefficient at 532 nm, the second one, a continuity test. The continuity test is

based on grouping bins along track into a two dimensional space and a new bin

is considered cloudy if at least half of pixels in the space are considered cloudy

by the first criterion.

The improved cloud mask from CALIPSO better classifies the clouds as

compared to the vertical feature mask (VFM) from original CALIPSO datasets;

especially for low clouds over dust. The new CALIPSO-CloudSat cloud mask

improves the CloudSat mask, especially for cloud tops below 720 m above the

ground (Hagihara et al., 2010).

CA-Ctype includes phase and type of cloud particle. Yoshida et al. (2010)

developed a method for discriminating cloud particle type based on the

relationship between depolarization ratio and ratio of backscattering coefficient

between two consecutive layers (X). It is based on the fact that liquid drops

have values of depolarization ratio as higher as for the case of ice drops, but

differences are found for X, because it is proportional to the first layer optical

depth. Discrimination based on temperature is also considered. The following

cloud particles types are considered: 3-D ice, 2-D plate, warm water and

supercooled water.

Cloud microphysics properties such as re for ice clouds and IWC are

included in CSCA-Micro product. The combinations of radar-lidar for retrieving

microphysics properties are more accurate than only using Radar (Okamoto et

al., 2010). Thus, re and IWC are retrieved by using lookup-tables where radar

reflectivity (Ze), depolarization ratio and attenuated backscattering coefficient

are the input variables. In the retrieval, specular reflection is taking into account

improving the accuracy of the method for cases with presence of 2-D plates

(Okamoto et al., 2010). The method is also applied for only radar or lidar cloud

mask, but input variables that cannot be measured are inferred from

relationships previously computed in the overlap radar-lidar layers in the same

profile.

The A-train trajectory does not overpass the campus USP (23.56° S,

46.73° W). The nearest overpass occurs at a minimum distance of 25 km with

34

the maximum frequency of overpasses near 75 km. Therefore, data from

overpasses distant at maximum 100 km from the ground site were selected.

3.2. Ground based measurements

At IAG meteorological station, far away 15 km to the Campus USP,

visual observations are made along with variables of weather conditions

measurements. Moreover, on the top of the Pelletron building (30 m height

above ground surface) at the Instituto de Fisica da Universidade de São Paulo

(IF-USP) in the Campus USP, three instruments are installed: sun-photometer

from CIMEL, model CE318, Multifilter Rotating Shadowband Radiometer

(MFRSR) from Yes Inc., and a broadband pyranometer from Kipp & Zonen,

model CM21. Furthermore, nearby at Center for Lasers and Applications (CLA)

of the Nuclear and Energy Research Institute, a total sky camera and Raman

lidar are also running continuously.

3.2.1. Visual observations of clouds1

At IAG meteorological station (23,65° S, 46,61° W, 800 m a.s.l) with a

since 1957, clouds are reported by human visual observations every hour from

7:00 Local Time (LT) up to 0:00 LT. A total of 12 cloud types are reported:

fractocumulus (Fc) and fractostratus (Fs), along with 10 main cloud types

defined by the World Meteorological Organization (WMO). In addition to cloud

type, the cloud amount (cloud cover) at each height level (low, middle and high)

is also reported (in tenths). This cloud amount is reported from a surface

reference point; and the sum of the cloud amount reported at each level

reaches a maximum of 10 tenths.

3.2.2. Pyranometer

The pyranometer measures G at the surface, integrated from 305 nm up

to 2800 nm, with a time resolution of 1 minute. It is provided with a thermal

detector where energy is absorbed and converted to voltage. Measured units

1 This section is part of Datasets and Methods of a paper submitted to the International

Journal of Climatology, ID=JOC-18-0083, for possible publication.

35

are converted to W m-2 using a calibration factor. It is calibrated about every two

years at LIM (Laboratório de Instrumentos Meteorológicos from Instituto de

Pesquisas Espaciais - INPE). Database of pyranometer measurements spans

from January 2016 until September 2017.

3.2.3. Multifilter Rotating Shadowband Radiometer (MFRSR)

Aside the pyranometer, two MFRSRs, numbered 345 and 368, make,

simultaneous measurements every 1 minute. The MFRSR measures 𝐺𝜆, 𝑆𝑆𝜆,

and 𝐷 ↓𝜆, at 5 narrowband channels centered at wavelengths: 415 nm, 670 nm,

940 nm, 870 nm, 1036 nm. It has an automatic band that blocks the sun

allowing to measure 𝐷 ↓𝜆. The 𝑆𝑆𝜆 is retrived from the difference between 𝐺𝜆

and 𝐷 ↓𝜆 using the CSZA as shown in the equation 17. In addition, the

instrument has a broadband silicon detector which measures global and diffuse

irradiance in the spectral range from 300 to 1100 nm.

The MFRSR 345 narrow band channels are calibrated every two years;

thus, spectral corresponding values at the top of the atmosphere (TOA) are

estimated. A total of 118523 measurements were made with this instrument,

spanning only 6 months of data from September 2016 up to February 2017.

Data from instrument 368 were employed for radiative computations and cloud

properties retrievals, since it was running from June 2012 until September 2017.

There are some measurement gaps for MFRSR 368: December 2012,

January 2014, from May to September 2014, November 2014, January 2015

and from April to July 2015. To calibrate MFRSR 368, coincident measurements

with MFRSR 345 were used, as will be detailed in section 3.5.2.

3.2.4. AERONET sun-photometer

Sun-photometer CIMEL, model CE318, is an automatic sun/sky spectral

radiometer with 1.2 ° field of view (FOV) which measures direct sun and sky

radiance in spectral narrow bands. The instrument placed in the site of USP

campus USP (23.56 °S, 46.73 °W) has 8 spectral bands centered at 340 nm,

380 nm, 440 nm, 500 nm, 675 nm, 870 nm, 940 nm, and 1020 nm. Direct sun

measurements are carried out in the eight bands. AOD (except at 940 nm) and

36

precipitable water vapor (at 940 nm) are retrieved from inversion methods using

spectral direct sun irradiances applying the Beer-Lambert-Bouguer law. It

returns to NASA Goddard Space Flight Center facility every two years for

calibration purposes.

Sky radiance measurements are employed for retrieving other properties

of aerosols e.g. size distributions, asymmetry parameter and single scattering

albedo (Dubovik and King, 2000). The surface albedo is also obtained by

combination of sky radiance and TOA measurements made by satellites

(Sinyuk et al., 2007). Spectral surface albedo is given at 440 nm, 670 nm, 870

nm and 1020 nm using Moderate Resolution Imaging Spectroradiometer

(MODIS) data, at or near the channels the of sun-photometer.

COD is retrieved from radiance measurements made at zenith at 440 nm

and 870 nm, in conditions where direct sun measurements are impossible to

make due to obstruction of sun disk by clouds (Chiu et al., 2010). The

methodology is better detailed in section 2.4.3. Since November of 2000 sun-

photometer is working at the site retrieving aerosol properties. COD began to be

retrieved in May 2016. Measurements of sun-photometer are made each 15

minutes, but in some cases around sunrise it is reduced to 5 minutes.

For AOD, there are 3 quality levels: level 1.0 which is not cloud screened,

1.5 for cloud screened and 2.0 for cloud screened and post calibration. In the

case of COD, there are two levels: 1.0 and 1.5. For COD, at level 1.0, the

retrieval is made by using climatological values of surface spectral reflectance

and for COD level 1.5, the spectral surface reflectance is obtained from MODIS

measurements. COD are retrieved every 15 min if clouds are present, while

each measurement takes about 1.5 min.

For this research, the following variables from sun-photometry were

used: spectral AOD, Ångström exponent, SSA, g, surface albedo, COD and

precipitable water vapor. In addition, ancillary information on NO2 and O3

column integrated data was also employed. Data were downloaded from the

site: https://aeronet.gsfc.nasa.gov/ .

3.2.5. Sky camera

37

At Center for Lasers and Applications (CLA) of the Nuclear and Energy

Research Institute (IPEN) (Latitude: -23.56°: Longitude: -46.74°, 740 m above

sea level), total sky camera J1006 is working since August 2016.

The camera is a weatherproof automatic system that captures cloud

cover with a fisheye lens. It has a resolution of 4M pixel and includes a

processor to perform digital images. Moreover, a powerful “Find clouds”

software allows to classifying the clear sky, total hemispheric cloud cover,

optically thick and thin clouds, direct sun detection and cloud cover of the free

and above artificial horizon.

In this research, regular pictures each 5 min were used from August

2016 to September 2017. Images can be obtained in the website:

http://gescon.ipen.br/leal/10-1.Measurements-2018.html .

3.2.6. Lidar

In the CLA of IPEN, since 2001, there is a multiwavelength Raman Lidar

installed. The fundamental emission is made at 1064 nm and additional

emissions at 532 and 355 nm are made using second and third harmonic

generators. The detection and spectral selection is made in the following

wavelengths: 1064 nm, 532 nm and 355 nm, 607 nm, 387 nm, 408 nm (da Silva

et al., 2017).

In this research, backscattering coefficient profiles of measurements at

532 nm were employed for around two days for each month of clear sky

conditions observed in 2015. A fixed value 60 of aerosol lidar ratio is assumed

and the profile defined between 300 and 6000 m above ground. Aerosol

extinction coefficient profiles from the system were employed.

3.3. Methodology of cloud climatology from visual

observation2

Cloud amount reported from visual observations per height level does not

represent the actual cloud cover for mid or high clouds when they are present

2 This section is part of Datasets and Methods of a paper submitted to the International

Journal of Climatology, ID=JOC-18-0083, for possible publication

38

with lower clouds. In order to infer the amount of clouds when they are hidden

by lower clouds or mid and lower clouds, a random overlap assumption (Tian

and Curry, 1989) is applied and shown in equation 34. The terms i and i-1 are

the current level and the level below it, respectively, N is the cloud cover or

cloud amount in the respective level. When high clouds are present with mid

and low clouds, cloud cover of the underlying level is the sum of cloud cover of

low and middle level clouds reported by the observer.

The real cloud cover NiR can be estimated with a maximum of 7/10 of the sky

obscured by underlying level clouds. With the presence of a Cumulonimbus

cloud (Cb) this equation is not employed due to the weakness of the

relationship (Tian and Curry, 1989).

NiR =

Ni

(10−Ni−1)∙ 100 (34)

For this research, only daytime (7:00-18:00 LT) cloud data were analyzed

due to restrictions to observe clouds at night. Observations from 1957 up to

2016 were considered, spanning a total of 59 years.

Cloud occurrence frequency (fq) for each cloud type is computed as the

ratio between the number of cases of each cloud type reported at a given height

level and the total number of cloud observations in that level. For mid and high

clouds, the height level is observable and considered viable if there is at least 3

tenths of the sky free of clouds to make sure that clouds are observable. This

condition is very important for high level clouds, because in several occasions,

the presence of lower clouds does not allow to observe possible higher clouds,

which consequently are not reported, producing underestimated frequency

values.

Cloud amount when cloud is present (awp) is the mean of cloud cover

computed when a cloud type is reported alone at a particular height level. Mean

cloud amount (amt), is the mean cloud cover computed, and for each cloud type

can be estimated from frequency and awp (equation 35). Values of awp are

converted to percent values for better analysis.

amt(%) = fq ∙ awp(%) (35)

39

In addition to the frequency of occurrence, co-occurrence frequency

(frequency of occurrence of one cloud type given another cloud type) is

computed. This variable symbolizes the dependence of one cloud from other

clouds. The most important cloud combinations by observed frequency were

also specified.

Trend of cloud amount is represented as the slope of a linear fit of mean

annual cloud amount as function of year, with statistical significance computed

with the modified Mann-Kendall Test (Hamed and Rao, 1998) at a significance

level of 5 %. Trend value is reported as %/ decade. Special emphasis is made

for diurnal cycle and it is analyzed in depth, computing phase and amplitude.

The phase was computed as the hour of maximum cloud cover and the

amplitude as the difference between maximum and mean cloud cover

computed for each day. Also, mean diurnal cycles for every 15 days along the

year were computed.

3.4. Radiative transfer model (uvspec from LibRadtran)

LibRadtran is a library containing programs and routines of radiative transfer

in shortwave and longwave spectra. The radiative transfer program uvspec, in

the shortwave spectrum, with 1-D configuration of LibRadtran version 2.0, was

employed in this study to compute downward solar irradiance reaching the

surface. The discrete ordinate method (disort) with 6 streams was selected as

the radiative transfer solver of the program (Stamnes et al., 1988).

The REPTRAN band parameterization (Gasteiger et al., 2014) with spectral

resolutions of 15 cm-1 was the adopted molecular absorption parameterization.

Trace gases, pressure and temperature profiles from the standard tropical

atmosphere (Anderson et al., 1986), where the atmosphere is divided in 50

levels from surface up to 120 km, were chosen. O3 and NO2 were modified from

the initial profile by using monthly mean values of total column concentrations.

O3 total column concentration given in Dobson units (DU) was obtained from

Total Ozone Mapping Spectrometer (TOMS) (McPeters et al., 1998) and

downloaded from: https://ozoneaq.gsfc.nasa.gov/. NO2 total concentrations in

molec/cm2 (Boersma et al., 2004) were obtained from ESA Scanning Imaging

Absorption Spectrometer for Atmospheric Chartography (SCIAMACHY), at:

40

http://www.temis.nl/airpollution/no2.html. These values are reported as output of

AERONET products. For computing broadband irradiance, the solar spectrum

at the top of atmosphere (TOA) of (Gueymard, 2004) with spectral resolution of

1 nm was used. For computing in narrow spectral range a better resolution of

0.1 nm from Kurucz (1994) was employed.

The spectral range selected for computing broadband irradiances is within

the spectral range of the pyranometer (305 nm up to 2800 nm). In narrow

bands, computations for MFRSR 345 were performed at 400.3 nm - 421.3 nm

and for MFRSR 368 at 404.3 nm-424.2 nm. Spectral response function was

taking into account for MFRSRs. Height above sea level was set up as 860 m

because of height of top of Pelletron building. For instantaneous or specific

cases, the day of the year and CSZA were also changed.

Column integrated water vapor included in the model was retrieved from

precipitable water vapor given by AERONET from direct sun measurements at

940 nm. This variable was included for computing clear sky irradiances. The

output of the model can be modified to irradiance units. The total transmittance

is calculated using equation 19.

3.4.1. Surface albedo

Surface albedo in the spectral range 280 nm-4000 nm was interpolated

using the relationship between the mean spectral surface albedos from

AERONET at the site and spectral reflectance of different typical surfaces of the

region, selected from the database of surfaces given by Baldridge et al. (2009),

e. g. asphalts, trees, concrete and grass.

For each surface type, in wavelengths where albedo is not reported by

AERONET, the ratio between surface reflectance at one wavelength and

surface reflectance in the nearest wavelength from AERONET was calculated.

Thus, using the mean of spectral surface albedos given by AERONET and the

mean of surface ratios at any wavelength, albedo at wavelengths not given by

AERONET were retrieved. The distribution of estimated mean spectral surface

albedo is shown in Figure 3. The surface is considered as lambertian in the

model.

41

3.4.2. Aerosols

Vertical profiles of aerosol particles were built using extinction coefficient

computed from backscattering coefficient at 532 nm measured by ground-based

Lidar. For each clear sky day, morning and afternoon mean profiles of extinction

coefficient were normalized with respect to the mean AOD computed by

integrating the extinction coefficient in the column, hereinafter called AODlidar.

Owing to values of extinction coefficient were retrieved only above 300

m, it was essential to know what fraction of total AOD was below 300 m. In

order to compute the contribution of AOD below 300 m, mean AOD from sun-

photometer (AODAERONET) in the same time interval of LIDAR was subtracted

from the mean AODlidar as long as AODAERONET was higher than AODlidar. Then,

a mean ratio between AOD from the lowest layer of LIDAR relative to AOD

below 300 m was calculated. AOD below 300 m was included in the mean

profile using the former ratio.

Figure 3: Mean spectral surface albedo used as input in the uvspec radiative transfer model.

Due to the lack of clear annual trend of stratospheric AOD in the last

decade, a mean value of stratospheric AOD of 0.008 was taken into account in

the computations (Vernier et al.,2011).

In figure 4.a, normalized extinction coefficient profiles from LIDAR, for

each season and part of the day, are shown. Note the maximum near 300 m

and the seasonal and diurnal differences. In the mornings, as expected,

42

aerosols are more concentrated near surface as compared to the afternoon,

when loading of aerosol increases aloft. Remarkable increase is observed

above 2 km due to the advection of biomass burning aerosols in spring (SON)

(de Miranda et al., 2017). With the aim of including aerosols in the radiative

transfer model, the normalized AOD profiles (𝐴𝑂𝐷𝑁𝑂𝑅𝑀(𝑍)) below 8 km

including AOD below 300 m were grouped in two seasons: typical: when

profiles of the spring are not included, characterizing mean local contributions

and spring (SON): mean profile for spring season, with advection of biomass

burning plumes. Profiles were also classified as morning and afternoon.

Figure 4: Mean profiles of normalized aerosol extinction coefficient for each season and time of the day (a) and the normalized profile of AOD built by synergy between LIDAR and sun-photometer for normal or typical conditions and for conditions with biomass burning plume advection in spring (SON) (b).

The spectral AOD profile included in the uvspec model was constructed

using the profile of normalized AOD and values of AOD retrieved by AERONET

(equation 36).

𝐴𝑂𝐷 (𝜆, 𝑍) = 𝐴𝑂𝐷𝐴𝐸𝑅𝑂𝑁𝐸𝑇(𝜆) ∙ 𝐴𝑂𝐷𝑁𝑂𝑅𝑀(𝑍) (36)

Narrowband irradiance around 415 nm was computed using AOD415,

estimated with AOD440 and the empirical formula of Ångström (equation 37),

using the exponent (𝛼) determined with the 380 nm and 440 nm AOD values

retrieved by AERONET.

43

For broadband computations, AOD at 500 nm was employed and

spectral dependence was included using the Ångström exponent 𝛼 and

Ångström’s turbidity coefficient (𝛽) (equation 37). This method has a good

accuracy as reported by Serrano et al. (2014). SSA mean values of 0.85 and

0.90 were assumed (de Almeida Castanho et al., 2008), for local conditions and

during spring time respectively.

𝐴𝑂𝐷(𝜆) = 𝛽𝜆−𝛼 (37)

Under cloudy conditions AOD is not retrieved by sun-photometry,

therefore four methods for estimating AOD at MFRSR 415 nm channel and

pyranometer were developed:

Time interpolation of AOD throughout the day;

Adjust climatological monthly diurnal cycle from at least one AOD

retrieved on the same day;

Adjust climatological monthly diurnal cycle of AOD from mean diurnal

AOD computed as near as at least 2 days;

Climatological monthly diurnal cycle of AOD.

Error in the estimation of AOD and contribution of error to COD retrieval by

each method will be analyzed and discussed in section 4.3.2.1.

3.4.3. Clouds

Ice and liquid clouds were included in the model and assumed as

homogenous layers. Cloud base and top, mean re and mean LWC/IWC are the

input variables for clouds. These inputs were assumed from microphysical and

geometric properties obtained from CALIPSO and CALIPSO-CloudSat as

explained before. In table 2, the default configurations for each cloud type are

summarized. Analysis of variability of cloud properties from satellite retrieval is

carried out in section 4.1.

The parametrization used for liquid clouds is given by Hu and Stamnes

(1993), where COD is obtained from re and LWC. They showed that the shape

of the cloud droplet size distribution does not affect the results (re represents

44

well the drop size distribution of clouds). This parametrization is as accurate

and faster in computation time than from Mie theory.

For ice clouds, the employed parametrization was from Key et al. (2002)

where optical properties of seven shapes of ice particles (habits) are

parametrized depending on re and IWC. According to the selected habit,

differences can be as higher as 15 %, especially between ice plate and the

other habits. The habit selected is the Rough Aggregated as an example of 3D

ice. In section 4.1 the selection of this habit is discussed.

Table 2: Default cloud configuration assumed in the radiative transfer code.

Cloud Top (km) Base (km) re (µm) LWC/IWC (g/m

3)

Ice 10.05 9.09 51.2 0.01 Liq. 1.60 0.86 11.6 0.23

3.5. Methodology of the radiative computations and COD

retrieval.

In the figure 5 the flow chart of the procedures for the retrieval of COD

and radiative effects computations are shown. The main datasets are the global

irradiance measured by the pyranometer and spectral diffuse and direct

irradiances at 415 nm and 870 nm measured by the MFRSR. Three steps

before the direct computations with LibRadtran are carried out. First, the 1 min

instantaneous calibrated measurements of MFRSR and Pyranometer are time-

collocated with precipitable water vapor and AOD from sun-photometer. The

second step is the definition of cloud scenarios by the Sky Definition method,

detailed in the next section. After the measurements are filtered, the mean

aerosol profiles according to the season and hour, in combination with the

coincident AOD from sun-photometer, are used as input in the radiative transfer

model. In addition, the mean microphysical properties retrieved from the

combination of CALIPSO-CloudSat, for ice and liquid clouds are inputted to the

model used in the retrieval of ECOD.

45

Figure 5: Flow chart of the main procedures carried out for the ECOD retrieval and computations of cloud radiative effects.

3.5.1. Sky definition

Ancillary information concerning cloud presence and obstruction of solar

disk are needed to classify the measurements made by pyranometer or

MFRSR. Because the time period of all sky-camera is limited (started in

September 2016), before this date, algorithms to define the presence of clouds

were required. Therefore three algorithms were developed for applying in

absence of sky camera observations: using visual observations of clouds,

spectral ratio of diffuse irradiance. In addition, algorithm of obstruction of solar

disk by clouds was defined. If the three algorithms were coincident, the

instantaneous measurements were classified. In the following the three

algorithms are detailed.

3.5.1.1. Sky definition by visual observations

Time coincidence with visual observations in a time window of 20

minutes before and after the time of observation was a criterion selected to

46

classify the sky by visual observations above the measurement site. Despite the

distance from the site is about 15 km in a straight line to the meteorological

station where visual observations are made, we considered that the spatial

variability in clouds between the two sites were not as high as to make

significant differences in cases when clear sky or total cloud cover were

reported.

Comparing with observations from the sky camera, with 6868 cases of

total overcast conditions, 99.68 % of coincidence was found with visual

observations from the meteorological station, when cloud cover greater or equal

9 tenths was observed. Below 8 tenths, a 100 % of coincidence were noted.

Thus, visual observation can be a useful ancillary data in dates when sky

camera observation was not available. For clear sky conditions, more than 95 %

of data observed as clear sky by the sky camera were reported also at the

meteorological station.

Classification of sky from visual observations is as follows:

Clear sky, if the maximum of cloud cover is 20 % with no presence of

high clouds.

Low/mid/high cloud partially overcast: only low/mid/high clouds with

cloud cover below 70 %.

Low/mid/high clouds total overcast: only low/mid/high clouds with cloud

cover of 100 %.

3.5.1.2. Direct sun criterion (solar disk obstruction)

In addition to cloud cover, information about obstruction of solar disk is

needed. With MFRSR measurements direct sun irradiance can be derived.

Then, comparisons of Tdir modeled for clear solar disk by uvspec and measured

Tdir at 415 nm were also used to classify the conditions of the solar disk. Min,

(2004) screened clouds by using direct sun irradiance when thin clouds were

present by applying threshold of Ångström exponent for aerosols. But it is only

limited to thin clouds and needs to compute spectral AOD with MFRSR which is

beyond the scope of this research.

47

The solar disk was considered obstructed by cirrus clouds when

modeled Tdir was 15% higher than the Tdir measured, otherwise it was

considered as a clear disk. This threshold was selected after the analysis of

uvspec error in computing Tdir at 415 nm for clear sky conditions (section 4.3.3).

The 15 % difference criterion works better for cirrus clouds with lower

COD values, but in the case of low clouds, with the presence of interstices, it is

not robust. Therefore to avoid misclassification in low clouds conditions,

differences between Tdir modeled for clear sun disk and Tdir measured must be

higher than 200 %. For mid-level clouds, 20 % of difference was considered.

These thresholds were stablished analyzing cases observed by the sky camera.

3.5.1.3. Spectral diffuse ratio (DR)

The presence of clouds can modify the spectral distribution of diffuse

radiance relative to ones of clear sky. The physical reason is that cloud drops

are larger than molecules and aerosols; therefore they present lower spectral

dependence in the solar extinction, i. e., compared to aerosols and molecules.

Thus a ratio between diffuse irradiance at 870 nm and 415 nm hereinafter

diffuse ratio (DR) can be a good indicator of the presence of clouds.

This ratio is based on the method applied by (Min et al., 2008) using the

diffuse transmittance to compute cloud fraction. Furthermore, LeBlanc et al.

(2015) employed the spectral ratio of radiance at zenith at 870 nm and 600 nm

measured at the surface. They showed the high sensitivity of the ratio at lower

CODs.

In figure 6, the variation of DR with CSZA, COD for cloudy and AOD500

for clear sky is evaluated using 1-D uvspec model. AOD500 was varied between

0.08 and 0.8. Between 0.08 and 0.3 the step was 0.03, while above 0.3 it was

equal to 0.1. COD for ice clouds spans from 0.03 up to 4, with steps of 0.5, and

for liquid clouds, it varied between 3 and 100, with steps of 5.As it can be

observed in the figure 6, DR varies between cloudy and clear skies. Below

AOD500 of 0.3, clear sky can be distinguished from sub visual cirrus with COD of

0.03. When single scattering dominates over multiple scattering, the sensitivity

with CSZA is observed. That can be noted for clear sky and for low COD of ice

48

clouds. When multiple scattering dominates DR is less sensitive to CSZA (low

clouds and higher COD of ice clouds). From the behavior observed for low

clouds is possible to clearly identify them over the clear sky cases. But

differences between ice and liquid clouds are not so obvious for thick cirrus

clouds.

Figure 6: Variations of diffuse ratio, at channels 870 nm and 415 nm, at surface level, with CSZA and AOD for clear sky, and COD of ice and liquid clouds computed from 1-D uvspec model of LibRadtran. The darker the color the higher the value of AOD500 or COD. COD for ice clouds spans from 0.03 up to 4 and, for liquid clouds, it varies between 3 and 100. AOD500 is varied between 0.08 and 0.8. Between 0.08 and 0.3 the step is 0.03, while above 0.3 is 0.1.

From the aforementioned analysis of DR, it is possible to make a criterion

to define clear and overcast sky, using spectral measurements of diffuse

irradiance from MFRSR. Therefore, after validation using real measurements

(section 4.3.4), the following criteria are used to define cloudy and clear sky:

Clear sky (clear) is defined when AOD500 is below 0.30, because above

0.30 can be overlapped with subvisual cirrus. Therefore depending on

CSZA for AOD below 0.3 clear sky cases can be defined by the

expression given in the equation 38.

𝐷𝑅 < 2.8 ∙ 10−2 ⋅ 𝐶𝑆𝑍𝐴−2 − 0.11 ∙ 𝐶𝑆𝑍𝐴 + 0.25 (38)

Total overcast sky, with high/mid-level clouds (mid or high overcast), is

defined by the equation 39, where coefficients a, b and c depend on

AOD500, to avoid overlap with possible clear sky.

49

𝐷𝑅 > 𝑎 ⋅ 𝐶𝑆𝑍𝐴−2 + 𝑏 ∙ 𝐶𝑆𝑍𝐴 + 𝑐 (39)

Total overcast sky with low cloud, mid or high clouds (low, mid or high

overcast) is classified if DR is above 0.6.

After the coincidence between the three methods (in the Sky definition),

before September 2016, 110727 cases for MFRSR 368 were classified

according to cloud scenarios shown in table 3. A total of 59592 of total overcast

sky cases, including low and high clouds, were employed to retrieve ECOD. For

pyranometer measurements, 26853 cases were classified.

In Table 3, a resume of sky definition algorithms to define cloud scenario

are shown. For the ECOD retrieval, cloud scenarios LW_0 and H_0 were

employed. For cloud radiative effects all the scenarios were used. Note, that

after September of 2016 only definition of visual observation from sky camera

pictures and direct sun criterion are employed.

Table 3: Sky definition algorithm for the clouds scenarios defined in this research.

Sky definition Cloud scenario Visual Direct sun DR

Low clouds total overcast obstructed Low overcast LW_0

Low clouds partially overcast obstructed Not clear LWBK_0

Low clouds partially overcast clear Not clear LWBK_1

High clouds total overcast obstructed High overcast H_0

High clouds partially overcast obstructed Not clear HBK_0

High clouds partially overcast clear Not clear HBK_1

Mid clouds total overcast obstructed Mid overcast MID_0

Mid clouds partially overcast clear Not clear MIDBK_1

3.5.2. Calibration of MFRSR 368

Transmittance from MFRSR 368 at 415 nm is the main input variable to

be used for retrieving the effective cloud optical depth. This instrument was

calibrated using indirect calibration (Holben et al., 1998). The idea is to transfer

the calibration of MFRSR 345 to MFRSR 368 using coincident measurements

under clear sky conditions. Cases of clear sky were selected from sky camera

after September of 2016. An inverse method from Lambert-Beer equation was

50

applied, using the AOD at 415 nm computed from MFRSR 345. Retrieval of

AOD415 from MFRSR was carried out by the same method applied by Sayão,

(2008).

The aim of the calibration procedure is to obtain the equivalent solar

irradiance at 1 astronomical unit (AU) (𝑆𝑆0) at TOA (equation 40). Where 𝑆𝑆𝑠 is

the direct sun irradiance measured at surface, 𝑘 is the correction factor of Earth-

Sun distance, 𝑚 is the relative optical air mass for Rayleigh scattering,

𝑀𝑂𝐷 𝑎𝑛𝑑 𝑁𝑂𝐷 are the molecular optical depth due to Rayleigh scattering and

nitrogen dioxide, respectively. 𝐴𝑂𝐷415345 is the aerosol optical depth at 415 nm

retrieved from MFRSR 345.

𝑆𝑆0 = 𝑘−1𝑆𝑆𝑠 exp[𝑚(𝐴𝑂𝐷415345 + 𝑀𝑂𝐷 + 𝑁𝑂𝐷)] (40)

For every 1 min of clear sky data, 𝑆𝑆0 was retrieved and time

interpolations were made for measurements with no clear sky conditions. For

cases when time interpolation was impossible to make, the nearest neighbor

interpolation was carried out. After calibration, T was computed using equation

18.

3.5.3. The retrieval of effective cloud optical depth

Since cloud optical depth was obtained from hemispherical

measurements under total overcast conditions, the term Effective Cloud Optical

Depth (ECOD) can be adopted. The retrieval of ECOD was carried out by

inversion method using 𝑇𝑡 at 415 nm computed from measurements of MFRSR

368. Transmittance was selected because errors associated to the calibration

accuracy were reduced. At around 415 nm, surface albedo is low, thus multiple

scattering between surface and cloud base is minimized. In addition, at this

wavelength there is only weak NO2 absorption and O3 absorption is negligible.

Furthermore, there is no spectral variability of COD at these wavelengths (Min

and Harrison, 1996). As the retrieval was applied for total overcast conditions,

and the model did not allow the inclusion of clouds with mixed phase, the

retrieval method was applied for either low (warm phase) or high clouds (ice

phase).

51

Visual observation with sky camera helped classifying total overcast

cases after September of 2016. Before this date, the sky definition criterion

(section 3.5.1) was applied. Cases under precipitation were discarded using the

report of precipitation made at the meteorological station.

Based on the method applied by Salgueiro et al. (2016), but including the

aerosol effect, lookup-tables (LT) for each cloud type (ice or liquid) were built

using the 1-D uvspec where output is 𝑇415.The input variables of each LT are:

CSZA, AOD415, COD and aerosol profile. The aerosol profile (figure 4.b) was

selected depending on season and time of the day. The sensitive study is

carried out in section 4.3.1, and the selection of input variables is discussed.

The procedure to retrieve COD is discussed as follows:

First, the aerosol profile is selected, afterward 2D cubic interpolation of

𝑇415 depending on CSZA and AOD415 for each COD step is carried out,

i.e. 𝑇415 =f(AOD415, CSZA). Therefore, 𝑇415 is estimated from the measured

AOD415 and CSZA for every COD step. Subsequently, linear interpolation of

estimated 𝑇415 as function of COD is computed. Using the linear interpolation,

ECOD is retrieved as the minimum difference between 𝑇415 estimated and 𝑇415

measured by iteration along the COD values.

In the configurations of each LT, AOD415 varied between 0 and 1.5 and

CSZA between 0.1 and 1 with steps of 0.05. For liquid clouds, COD varied from

0 up to 400 with step of 1. In the case of ice clouds, COD varied from 0 up to

40, with steps between 0 and 4 of 0.25 and 1 above 4.

3.5.3.1. Error in COD retrieval

The error propagation for COD retrieval is expressed in equation 41, as

the root square of the sum of error squared of each variable i. The relative error

is computed as the ratio of each error and the COD value retrieved.

𝛿𝐶𝑂𝐷 = [∑ (𝜕𝐶𝑂𝐷

𝜕𝑖𝛿𝑖)

2

𝑖 ]

1

2 (41)

Applying the methodology proposed by Serrano et al. (2014) (equations

42-44) the contribution of each variable to the COD error is retrieved. The

52

variation of T due to the error of variable i assuming the other variables constant

is 𝛿𝑇𝑖. In addition, the variation of T due to the possible error induced in COD

(𝛿𝐶𝑂𝐷𝑖) is 𝛿𝑇𝑐𝑜𝑑. Differences between 𝛿𝑇𝑖 and 𝛿𝑇𝑐𝑜𝑑 are computed by iteration

of 𝛿𝐶𝑂𝐷𝑖. When the difference between 𝛿𝑇𝑖 and 𝛿𝑇𝑐𝑜𝑑 is the minimum, 𝛿𝐶𝑂𝐷𝑖 is

obtained. The computation of T is carried out using the uvspec considering

CSZA equal 0.86.

𝜕𝐶𝑂𝐷

𝜕𝑖𝛿𝑖 ≈ 𝛿𝐶𝑂𝐷𝑖 (42)

𝛿𝑇𝑖 =1

2|𝑇(𝑖 + 𝛿𝑖, 𝐶𝑂𝐷) − 𝑇(𝑖 − 𝛿𝑖, 𝐶𝑂𝐷)| (43)

𝛿𝑇𝑐𝑜𝑑 =1

2|𝑇(𝑖, 𝐶𝑂𝐷 + 𝛿𝐶𝑂𝐷𝑖) − 𝑇(𝑖, 𝐶𝑂𝐷 − 𝛿𝐶𝑂𝐷𝑖)| (44)

Because of the transmittance error is directly related to the error

associated to the calibration procedure, 2 % of error is assumed. For single

scattering albedo, the error of retrieval from sun-photometer is about 0.05

(Dubovik et al., 2000), since the error decreases to 0.03 for AOD440 above 0.20

and around 0.05-0.07 for AOD below 0.20. SSA mean values of 0.85 and 0.90

were assumed (de Almeida Castanho et al., 2008), for local conditions and

during spring time, when biomass burning plumes are transported to the city,

respectively. The error of assuming mean value of SSA is the standard

deviation of SSA obtained from sun-photometer at the site. Therefore the total

error of SSA (𝛿𝑆𝑆𝐴), including the error of the retrieval is 0.19.

For AOD, the error assumed is the sum of error of AOD measurement

(0.02) (Serrano et al., 2014) and the maximum error of inferring AOD, from

mean monthly diurnal cycle at 415 nm (0.11) (Table 6). The standard deviation

of samples of cloud properties was assumed as the cloud properties errors. For

ice clouds error of re (𝛿𝑟𝑒) is about 40 µm and for low clouds, 2 µm. For cloud

base error (𝛿𝐶𝐵) it is computed as 2.8 km and 2.7 km for ice and liquid clouds,

respectively. The results of error propagation are presented in section 4.3.2.

3.5.4. Instantaneous cloud effects on solar radiation

The instantaneous cloud effect on solar radiation was evaluated by CRE

and NCRE (equations 26 and 29 of chapter 2). Irradiances measured every 1

53

min from pyranometer were selected as cloudy conditions, while irradiance for

clear sky was estimated using uvspec model.

Default configuration, explained in the topic 3.4, was used to calculate

irradiance under clear sky conditions. As will be discussed in the next chapter

(Results), in general, the model overestimated clear sky irradiances compared

to the measurements. The maximum relative difference of 15% and 5 % was

observed for overestimation and underestimation, respectively. To avoid

erroneous estimation of the cloud effect on solar radiation, when the relative

difference between modeled and measured irradiance was higher than – 5 %

and less than 15 %, the case was discarded.

For solar disk blocked by clouds, the following cases (according to table

3) were analyzed: completely or total overcast sky with low clouds (LW_0),

partially overcast with low clouds (LWBK_0), total overcast sky due to high

clouds (H_0), partially overcast sky with high clouds (HBK_0) and total overcast

sky due to mid clouds (MID_0). When the solar disk was clear, the following

cases were computed: low clouds (LWBK_1), high clouds (HBK_1) and mid

clouds (MIDBK_1).

54

4. Results

4.1. Cloud characteristics from CALIPSO-CloudSat

Seasonal profiles of frequency of occurrence of clouds for nighttime

(01:00 LT) and afternoon (14:00 LT) overpasses are shown in figure 7. There

were more than of 2000 profiles measured at night and more than 7000

measured in daytime for each season. Seasonal and day/night differences are

clearly observed.

Figure 7: Profiles of frequency of occurrence of clouds per season at nighttime (a) and daytime (b), computed from cloud mask generated by merged CALIPSO-CloudSat data.

Higher frequencies of clouds were observed in summer, whose vertical

development could reach 17 km. In winter, lower frequencies were found and

clouds could reach 15 km. The contrast between daytime and nighttime profiles

is observed. In daytime profiles the maximum frequency, around 30 %, takes

place near 2500 m in summer. As a consequence of the diurnal heating, cloud

frequency increased at lower layers during daytime. At night, cloud frequency

decreased a little below 2 km, and increased at mid (7.5 km) and high layers

(12 km) with 35 % of frequency in summer, as a consequence of the remaining

convective activity in the afternoon. At night, convection tends to disappear and

the water vapor supply from surface decreases, low clouds begin to evaporate

until radiative cooling starts the formation of stratiform low clouds (Wood, 2012).

55

In mid and high layers, clouds remain due to the weak circulation and the

radiative cooling which maintains or increases the clouds.

At daytime, from summer to winter, the decrease of cloud frequency is

observed more pronounced at lower layers from about 32 % to 14 %. At

nighttime higher differences are found in higher layers from 35 % to 5 %. In

winter, the maximum frequency is observed in lower layers with 15 % at

nighttime and minimum in higher layers. This maximum in lower layers at

nighttime is a consequence of the increase of radiative cooling.

In the figure 8, profiles of cloud frequency as function of the distance

from the coast are presented. In the computations, measurements at nighttime

and daytime were used together. Variability is also observed as the distance

from the coast line increases.

Figure 8: Frequency of presence of clouds relative to the distance from the coast line.

Note near the coast up to 20 km the convection activity is expected to be

less strong compared to inland. With a distance below 20 km far from the coast

line, the maximum frequency at lower layers (2.5 km a.s.l) with 26 % is

observed. From the edge of Serra do Mar mountain barrier (from 20 km) an

increase of cloudiness in mid and high layers is observed with maximum

56

frequency near 50 km from the coast, with 24 % frequency of occurrence near

10 km of height.

While the distance to the coast increased, low clouds frequency begin to

decrease and an increase of the height of clouds and frequency is observed.

Maximum of cloud frequency was found 100 km away from the coast and near

12 km height with 32 %. Undoubtedly, this behavior illustrates that convective

clouds are more frequent inland.

Geometrical properties, cloud base height and cloud thickness for ice

and liquid clouds for each season are shown in figure 9. Moreover, the number

of identified clouds is also shown above of each boxplots. In the case of cloud

base of warm clouds a slight variability is observed, with similar values for the

median in all seasons, but, in winter, heights of cloud base are more

concentrated around the median with the smallest interquartile range. However,

hypothesis test of Kruskal-Wallis (Kruskal and Wallis, 1952) does not show

differences between seasonal samples. Cloud base of liquid clouds at the

region presents the median value of 860 m.

Figure 9: Seasonal variations of geometrical properties of clouds: the height of base and thickness of liquid clouds (a, b) and ice clouds (c, d). The number of clouds identified is also shown in the upper border of each figure.

As expected, in summer, warm clouds are deeper with median of 900 m.

In winter and spring, the median of cloud thickness is around 600 m. For ice

clouds differences are larger, cloud base is higher in summer with a median of

57

11 km, and lower in winter (9.5 km). According to Kruskal-Wallis H-test,

differences are statistically significant. The thicker of ice clouds in winter,

probably has to do with the lower cloud base.

The seasonal variability of microphysical properties for ice and liquid

clouds is shown in figure 10. Liquid clouds show small sizes as compared to ice

clouds. Seasonal differences with statistical significance between re of liquid

clouds re confirmed by the Kruskal-Wallis H-test. The highest value of the

median occurs in winter with 11.8 µm, because in this season clouds are more

frequently originated from interaction of cold air mass outbreaks and sea, as a

result of cold migratory anticyclone systems.

Figure 10: Microphysics properties re and LWC or IWC for liquid (a, b) and ice clouds (c, d). For liquid clouds, the products of CloudSat are employed and for ice cloud, CSCA-Micro product was considered. The number of cases is placed in the upper border of each figure.

Similar to re, LWC has differences with the highest median in winter, but

with the minimum observed in spring. For ice clouds, the seasonal differences

are smaller with many outliers for effective radius, above 200 µm, while the

median values are around 51 µm. The general median value, considering all

data available, is equal 51.23 µm. IWC also presented many outliers with values

higher than 0.5 g /m3, with some values comparable to the maximum LWC

observed for low clouds, although the median presented little seasonal

differences with values around 0.01 g /m3.

58

To try to know the more frequent particle habit in the region, the product

of JAXA CA-Ctype for cloud particles was used. The frequency of occurrence of

cloud particle habit per height level for each season is presented in figure 11.

Three types of habit were analyzed: 3D-ice (3dic), 2D-plate (2dpl) and

mixture of 3D-ice and 2D-plate (mx3i2p). The predominance of three

dimensional ices is observed specially above 10 km where almost 100 % of ice

cloud layers are composed of it. In addition, 2D-plate habit is more frequent in

mid-levels particularly between 6 and 8 km with higher frequency in autumn and

winter probably due to the weaker updrafts (Heymsfield et al., 2017) as

compared to summer and spring. At 4 km to 8 km there is higher variability of

habit being impossible to define a fixed classification.

Figure 11: Profiles of frequency of occurrence of ice cloud habit for each season (a) and frequency of occurrence of cloud layer totally composed by a given cloud particle habit.

The frequency of occurrence of cloud habit is presented in figure 11 b. In

the region, clouds are predominantly composed by 3D ice particles with 95 % of

frequency, followed by the mixture of 3D with 2D plates with 4 % frequency and,

in the case of only 2D plates; they were identified in 1 % of data only. After this

analysis, the type of habit chosen as input in the uvspec of LibRadtran code for

ice clouds was the one with 3D geometry (as example 'Rough Aggregated').

Variations of microphysics properties inside the ice clouds respect to the

height above the cloud base are shown in figure 12.a-b. Median values and

interquartile ranges are presented. As it can be seen, ice clouds are far from

and idealized homogenous cloud. The decrease of cloud particle sizes and ice

59

water content is observed with the height above cloud base. The re can

decrease from 60 µm up to 40 µm near cloud top. But for IWC the reduction is

higher from near 10 mg m-3 until 2 mg m-3 near the top of cloud. This pattern is

a response to the decrease of water vapor availability as the temperature

decreases with the height.

Variations of mean microphysical properties of all cases of clouds

respect the height of cloud base are shown in figures 12 c-d. Mean re decreases

as cloud base is higher from 6 km (102 µm) up to 12 km (40 µm). However,

near 13.5 km a slight increase is observed. For IWC this behavior is also

observed with a higher increase near 13.5 km which can be attributed to cirrus

from remaining storms.

Figure 12: Variations of re and IWC with height above cloud base (a, b) and variations

of mean re and IWC of the cloud layer according to the height of cloud base (c,d).

4.2. Cloud climatology from visual observations3

4.2.1. Diurnal and annual cycles of cloud amount

3 This section is part of Results of the paper submitted to the International Journal of

Climatology, ID=JOC-18-0083, for possible publication.

60

Figure 13.a shows the mean diurnal cycle and respective variability of

clouds amount from visual observations. Figure 13.b-c gives indication of the

associated annual variability, i.e. how the hour of maximum and amplitude of

diurnal cycle vary along the year.

Figure 13: Mean diurnal cycle and standard deviation (a) computed from mean diurnal cycles calculated for every 15 days. Variability of the diurnal cycle for every 15 days computed as hour of maximum (b) and Amplitude (c).

The mean total cloud amount (clouds) decreases after sunrise until

about 11:00 LT (63 %), with a monotonically increase with the advance of the

day up to a maximum at sunset (74 %). This behavior is initially associated with

the possible increase of cloud amount in the end of the night, due to radiative

cooling. After sunrise, the radiative cooling is surpassed by solar heating,

breaking the inversion layer above cloud, and dissipating the cloud layer. The

posterior increase after 11:00 LT is due to convective development caused by

the solar heating, influence of the sea breeze or a synoptic system. If

convection is deep, formation of high and mid-level clouds are possible. High or

mid-level clouds, formed by convection, remain or increase in cloud amount at

night driven by radiative cooling (Heymsfield et al., 2017). Hence, mid and high

clouds present an inverse diurnal cycle of cloud amount as compared to solar

heating diurnal cycle (figure 13.a.).

61

This diurnal cycle of clouds is variable along the year (Figure 13.b-c),

with the maximum observed at sunset from September to March and around

sunrise from April to August. The decrease observed near sunrise (Figure 13.a)

takes longer in wintertime and lasts two hours more (until about noon)

compared to summer (not shown in figure). Undoubtedly, differences along the

year were due to differences of driven mechanism of cloud formation. In winter,

radiative cooling dominates, there is less deep convection and air humidity is

reduced, whereas in summer, deep convection is more frequent.

Low clouds drive the diurnal cycle of clouds as can be observed by the

similarity in their cycles. Low clouds has a significant influence of low stratiform

clouds, since they are the more abundant type of clouds in the location, and

only a pronounced increase in the diurnal cycle at noon and a slightly decrease

near 16:00 LT are due to cumuliform clouds.

Note the differences between low stratiform and cumulus clouds.

Stratiform clouds are less influenced by the solar cycle, rather the radiative

cooling effect is responsible for the formation at night and sunrise (Wood,

2012). When isolating low stratiform clouds (not shown), it is possible to

observe that the maximum cloud amount occurs for stratus clouds at sunrise

and for stratocumulus at sunset, thus, from Figure 13.b, maximum of diurnal

cycle probably can be influenced because stratus prevails in winter, while

stratocumulus dominates during the summer season. It is important to observe

that other mechanisms can lead to the formation of stratiform clouds, not only

radiative cooling, particularly in summer and spring, as will be discussed later.

This behavior was not observed when analyzing the synoptic reports (Eastman

and Warren, 2014) because of the underestimation of Sc when they are present

with Cb. Moreover in the region, there can be an important contribution of

cumulonimbus clouds by the possible influence of the sea breeze circulation,

synoptic systems as SACZ and cold fronts. The amplitude of stratiform clouds

was less variable along the year with maximum in autumn-winter (Figure 13.c),

and minimum in April. In this month, the increase of cloud amount at sunset is

less pronounced and the radiative cooling at sunrise is negligible, producing

lower amplitude in the diurnal cycle. In addition, in this month, cold fronts are

highly frequent (Morais et al., 2010).

62

Cumulus clouds, by contrast, present high dependence on the solar

heating diurnal cycle, with maximum near noon, throughout the year (Figure

13.b). Furthermore, this maximum occurs two hours later in winter compared to

summer. That is because, in winter, the maximum of cloud cover is found at

sunrise leading to delaying of surface heating by solar radiation, thus, the

increase of cloud amount by convection can start only in the afternoon. Note the

differences in amplitude of cumulus along the year, with the highest amplitude

in summer of about 15 % and minimum in winter, near 5 %.

The maximum amplitude, above 15 %, is also observed in summer for

mid-level clouds. Mid and cirriform clouds show comparable characteristics with

maximum in spring-summer at sunset, except on the first 15 days of October

when the maximum is observed at 7:00 LT. In autumn-winter, one observes a

maximum of mid clouds at sunrise but, for cirrus, the maximum can take place

near sunset. Note that, for cirrus, from January to March, the maximum is

observed in the morning.

Figure 14: Mean annual cycle computed for every 15 days. The vertical bars represent the inter-annual variability.

The annual cycle, computed every 15 days, is shown in Figure 14. Mean

values and standard deviation are presented. Observe the seasonal variability

for total cloud cover (clouds) with a minimum in early August. Maximum takes

place in summertime, between December and January. Low clouds show

63

minimum in the first part of August but, in contrast to total cloud cover, an

increase is observed between March and April due to the same increase

detected in this period for stratiform clouds. As mentioned before, the peak of

cold front systems crossing the region, as diagnosed by Morais et al. (2010),

can contribute to this increase.

Cirriform clouds presented less contrast between winter and summer,

with differences of about 15 %, with the maximum observed in spring, in late

October of 35 %. Mid-level clouds have a maximum of amt in the beginning of

January (40 %) and minimum on the first 15 days of August with 17 %.

Cumuliform clouds have an observable seasonal behavior with differences

between winter and summer of 18 %. The maximum is observed in the end of

December with 20 % and the minimum on the last 15 days of September (2%).

4.2.2. Inter-annual variability of cloud amount

The inter-annual variability of cloud amount is analyzed in this section.

Seasonal mean values are shown in Figure 15 for the six main cloud

categories. Seasonal variability and trends are observable especially for low

stratiform and mid-level clouds. Low clouds have an increasing trend due to the

stratiform variability. This increase is observable from the second half of 1980’s.

Total cloud amount (Clouds) has not a clear trend in any season, with

maxima of annual means occurring in summer. Absolute maximum and

minimum of total cloud amount is perceived in summer of 1975 (87.7 %) and

winter of 1963 (42.7 %), respectively. For mid-level clouds, a decrease after

1990 in all the seasons is evidently observed. The maximum occurred in

summer of 1965 (53.2 %) and minimum in winter of 2010 (1.9 %). Conversely,

cirriform clouds presented a slight increase especially in summer with maximum

observed in summer of 2010 (41.4 %) and minimum in winter of 1964 (7.4 %).

Furthermore, low clouds presented a maximum in spring of 2005 (70.1 %) and

minimum in winter of 1972 (25.2 %). In the same years, the maximum and

minimum of stratiform clouds were observed, with 71.7 % and 24 %,

respectively.

64

Figure 15: Mean values of cloud cover (amt) per year computed for every cloud type in summer (a,b ), autumn (c,d ), winter (e,f), spring (g,h).

Cumuliform clouds presented the maximum in summer of 2001 with an

absolute value of 24 % and minimum in winter of 1995 with 0.5 %.

Trend values computed as the slope of linear regression are shown in

Table 4. We estimated annual trends and for each season separately. Bold

values indicate trends that are statistically significant at 5 % confidence level.

These values were computed for the total time interval and for the first and last

30 years.

Table 4: Seasonal and annual trends (%/decade) computed for each cloud type, in three different time intervals. Statistical significance was computed by the modified Mann-Kendall Test at 5 % of confidence level and they are highlighted in bold.

Cloud Type

PERIOD TIME 1958-2016 1958-1988 1987-2016

DEF MAM JJA SON Total DEF MAM JJA SON Total DEF MAM JJA SON Total

Low 2.0 1.8 1.0 1.8 1,6 -0,1 0,5 -0,6 1,0 0,2 4,4 4,1 2,3 2,3 3,2

Mid -3.1 -2.2 -1.6 -2.7 -2,4 -0,3 2,2 0,5 -1,3 0,3 -7,3 -5,3 -3,1 -4,8 -5,2

Cirr 1.0 1.1 0.4 0.7 0,8 2,3 1,9 2,1 0,0 1,6 2,2 2,2 0,2 2,4 1,7

Str 3.7 3.5 2.0 3.0 3,1 5,9 5,7 2,1 5,0 4,8 1,5 2,2 1,3 0,8 1,3

Cu -0.1 -0.5 -0.5 -0.3 -0,3 0,1 0,3 -0,4 0,2 0,1 0,7 0,0 0,2 0,2 0,2

Clouds 0.2 0.8 0.1 0.4 0,4 0,8 2,3 1,2 0,3 1,1 0,1 1,2 0,3 0,4 0,6

From 1958 to 1988 an increasing trend is observed, but only with

statistical significance in all seasons for stratiform low clouds. For cirriform

clouds, statistical significance is observed in winter (2.1 %/decade) and for all

65

the year with 1.6 %/decade. For total cloud cover only in autumn an increase

was observed (2.3 %/decade).

After 1987 the increasing trend is higher for low clouds (3.2 %/decade)

due to a little increase with no statistical significance of Cu and stratiform low

clouds. In addition, it is noteworthy the decrease observed for mid-level clouds

in all seasons, with trend with annual mean values of about -5.2 %/decade. In

addition the increase of cirriform clouds of 1.7 % was noted in this period time.

For the total period of study, from 1958 to 2016 the decrease in mid-

level clouds amount in all the seasons with statistical significance is observed.

The maximum decrease of about -3 %/decade was found in summer. Cirriform

clouds showed a positive trend with significance in an annual basis (0.8

%/decade) and in autumn (1.1 %/decade). Cumuliform clouds presented a

decreasing trend with statistical significance, except for summer. But low clouds

increased in all seasons (above 1 %/decade) as a consequence of the highest

increasing trend observed for low stratiform clouds (above 2 %/decade) in all

seasons.

Comparing trends computed for a grid near the region for daytime data

by Eastman and Warren (2013) for the period 1971-2009, differences in low

cloud cover were found. While they reported a decreasing trend, it is clearly

noted an increasing trend of low clouds in this study. These differences are due

to stratiform clouds, because in this study a significant increasing trend was

observed for all season, whereas for Eastman and Warren (2013) a minor

increase was only observed in summer and autumn. Moreover, low cumuliform

cloud trends showed the best agreement between both studies, with reduction

in all seasons.

Trends of high and mid-level cloud almost agreed, presenting higher

decrease in our study. Differences in trends were only observed in summer for

mid-level clouds and spring for high clouds. For the total cloud cover trends,

similar results were found, except for spring and summer when positive trends

were observed in this study, but with no statistical significance.

66

Seasonal trends at distinct local time are shown in Table 5. It is clearly

observed a decrease for mid-level clouds during daytime hours in all seasons

with maximum decrease at sunrise. The absolute maximum occurred in spring

during sunrise (-3.8 %/decade). Low and stratiform cloud amounts were

increasing in the afternoon with maximum at sunset, except in winter, when

maximum trends occurred around noon, while cumuliform clouds showed

decreasing trend near noon, except during summer. Cirriform cloud amount

increased in sunrise and midday, but with statistical significance around noon

for autumn, spring and in the annual basis, and around sunrise in summer. For

total cloud cover, an increasing trend of 0.8 %/decade around noon and a

decreasing trend of -0.5 %/decade at sunrise were observed on an annual

basis. In autumn, cloud cover showed increase after midday and decrease at

sunrise.

Table 5: Trends (%/decade) of 6 six clouds types computed for sunrise (7 SLH), noon (13 SLH) and sunset (18 SLH). Statistical significance was computed by Mann- Kendall Test in 5 % and they are in bold.

Cloud Type

DJF MAM JJA SON Total

7 SLH

13 SLH

18 SLH

7 SLH

13 SLH

18 SLH

7 SLH

13 SLH

18 SLH

7 SLH

13 SLH

18 SLH

7 SLH

13 SLH

18 SLH

Low 0,4 2,6 3,4 -0,1 2,8 3,5 0,6 1,2 1,0 0,6 2,2 2,6 0,4 2,2 2,6

Mid -3,5 -3,1 -3,3 -3,3 -1,8 -3,1 -2,5 -1,4 -2,1 -3,8 -2,2 -2,7 -3,2 -2,2 -3,0

Cirr 1,3 0,7 -1,0 1,2 1,4 0,1 0,9 0,2 -0,2 0,6 1,1 0,6 1,1 0,8 -0,2

Str 0,6 6,2 4,2 0,0 6,5 3,7 0,7 2,8 1,3 0,7 4,6 2,9 0,5 4,9 3,0

Cu 0,1 0,4 0,2 0,1 -1,1 0,2 0,1 -1,0 0,0 0,2 -0,7 0,2 0,1 -0,6 0,1

Clouds -0,4 0,6 0,2 -0,7 1,7 1,2 -0,2 0,1 -0,4 -0,5 0,9 0,6 -0,5 0,8 0,4

4.2.3. Simultaneous occurrence and combinations of clouds

Following the methodology applied by Warren et al. (1985), in this

section, the simultaneous occurrence of cloud types is analyzed. Table 6 shows

cloud frequency of occurrence and cloud frequency of co-occurrence (cloud

type occurrence given another cloud type) for each season. In this case, Cu,

Sc, St, Cb, Ac, Ns and Cirriform clouds were employed. As can be seen, almost

all cloud types presented maximum frequency in summer except for stratus

clouds, whose maximum was observed in spring. In this season, the frequency

of cold front passage in the region is high, with 27 % of events of the year

(Morais et al., 2010).

67

Table 6: Frequency of occurrence and co-occurrence for clouds types for each season

Given

cloud

season Fq Frequency of occurrence due to given cloud type

Cu Sc St Cb Ac As Ns Ci

Cu

DJF 40.7 - 56.8 0.9 4.4 32.3 14.5 0.0 46.8

MAM 33.2 - 58.3 1.2 1.8 21.5 6.1 0.0 31.8

JJA 13.9 - 54.1 0.9 0.3 14.9 2.9 0.0 25.9

SON 22.7 - 55.9 0.9 1.6 21.4 7.7 0.0 34.9

Sc DJF 57.4 23.1 16.2 3.4 43.5 33.8 0.1 48.3

MAM 51.9 19.4 16.3 1.4 32.5 18.7 0.0 33.9

JJA 33.8 7.5 23.7 0.2 25.4 14.3 0.0 28.6

SON 49.2 12.7 24.5 1.0 32.8 25.0 0.0 36.1

St DJF 16.1 0.3 57.7 2.0 34.8 67.1 0.0 52.5

MAM 17.2 0.4 49.4 0.9 29.4 39.8 0.1 35.4

JJA 20.6 0.1 39.0 0.2 23.3 32.4 0.0 28.9

SON 25.4 0.2 47.6 0.6 30.3 54.8 0.1 40.3

Cb DJF 3.2 1.8 59.6 9.7 54.1 48.1 0.5 61.6

MAM 1.2 0.6 61.1 12.8 46.3 44.7 0.0 62.1

JJA 0.1 0.0 58.8 35.3 54.8 45.2 0.0

SON 0.9 0.4 52.9 15.3 48.4 42.4 0.5 67.2

Ac DJF 40.1 10.0 46.8 2.5 2.8 41.0 0.0 69.6

MAM 28.2 5.8 45.6 3.1 1.2 31.2 0.0 57.0

JJA 18.6 1.8 28.4 2.7 0.2 28.6 0.0 51.1

SON 27.6 3.9 38.7 3.4 1.1 41.2 0.0 61.8

As DJF 24.8 4.5 58.8 7.9 4.0 66.2 0.1 89.4

MAM 13.2 1.6 55.9 8.9 2.4 66.6 0.0 88.3

JJA 7.9 0.3 38.1 9.0 0.4 67.8 0.0 86.9

SON 17.2 1.4 47.2 9.8 1.5 65.9 0.1 89.6

Ns DJF 0.1 0.0 30.3 0.0 12.1 12.1 39.4

MAM 0.0 0.0

JJA 0.0 0.0

SON 0.0 0.0

Ci DJF 51.7 11.5 33.1 1.1 1.8 31.0 8.1 0.0

MAM 36.8 7.5 32.7 1.8 0.9 24.5 5.1 0.0

JJA 32.8 2.9 16.1 1.3 0.1 16.7 3.4 0.0

SON 38.3 5.6 26.4 1.5 0.8 21.9 5.5 0.0

Likewise, it is noticeable the Sc maximum in summer in contrast with the

global studies of Warren et al. (2007) and Wood (2012). However, it is important

to keep in mind that the Metropolitan Area of São Paulo (MASP) is located over

a plateau of 800m above sea level and close to the coast (about 50 km in

straight line).

68

Due to the parallel orientation of the mountain range to the coast line,

mountain-valley circulation contributes to the advance of sea breeze inland. In

addition, circulation generated by the urban heat island effect makes the sea

front to be stationary at MASP. This phenomenon influences the local

circulation and, acting isolated or with other systems, contributes to many cases

of floods in the region (Freitas et al., 2009). From the analysis of wind direction,

Perez and Silva Dias (2017) found that the sea breeze reaches the city earlier

in summer and spring (around 14:00 LT), compared to the other seasons.

In summer and spring the early arrival of the breeze is associated to the

higher contrast between the ocean and land surface temperatures. The

conditions are most favorable due to the increase of the land temperature and

the position of the South Atlantic high pressure. The breeze front can contribute

to the formation of convective clouds along with a composite of clouds such as

stratocumulus. Besides, possible advection of low stratiform clouds formed near

the coast and in the edge of the mountain barrier can occur. Note that more

than 50 % of Cb cases were in the presence of stratocumulus in all seasons.

Conversely, cumuliform clouds frequency, when Cb was present, was near null.

Dynamic processes associated with Cb can be another forcing mechanism for

stratocumulus development in the area. Downdraft outflow from Cb can lead to

the formation of other low cloud types, by means of formation of gust front

(Mahoney, 1988). This was not reported in previous studies using synoptic

reports, probably because of the constraint mentioned before related to the

presence of Cb and the way of synoptic observations were carried out.

It is also significant the higher dependence of mid-level and cirriform

clouds on cumulonimbus. Observe that cirrus is present with almost all cloud

types, with especially higher frequency with presence of As, Ac and Cb.

Likewise, note how Sc and Ac are related especially in summer, once Ac can

form from ascending layers of stratocumulus clouds.

For the analyzed period, 536 combinations of clouds were found; the top

10 by frequency of occurrence for each season are shown in the Figure 16. As

can be observed, there is an evident predominance of stratiform cloud

combinations except for Cu-Ci, Ci and Cu. The highest frequency was found for

69

Sc with maximum of about 16 % in spring. Maxima frequency are also observed

in spring for combinations with stratiform clouds; in summer for the combination

with cumuliform clouds; and, in winter, for Ci and Ac. Mean cloud cover when

the combination is present (awp) is shown in Figure 15.b. Observe the higher

values for St, Sc-St, Sc-Ac and Sc-Ac-St. Minima of awp were observed for Ci

and Cu with 50 % and 45 %, respectively.

The diurnal cycle of the first 8 combinations is shown in Figure 17.

Combinations of stratiform clouds show an inverse diurnal cycle compared to

combination with cumuliform clouds. At sunrise, the highest frequency for

stratus combination is observed. Also, Sc showed increase in the morning with

maximum near 9:00 LT. Observe the maximum frequency of cirrus-only at

10:00 LT, following the decrease observed for low cloud cover. Around noon

Sc, Sc-Cu and Cu were predominant. At sunset Sc has the highest frequency

followed by stratus clouds. Observe, in the figure 17.b, that cloud cover, when

combinations are present, is higher throughout the day for low stratiform

combinations, and the influence of the solar heating cycle on cloud cover of

cumuliform combinations is also noted.

Figure 16: Mean frequency of occurrence (fq) and sky cover when present (awp) for the12 main cloud combinations observed at each season.

70

Figure 17: Diurnal cycles of the 8 main combinations of clouds for frequency of occurrence (a) and amount when present (b).

4.3. Sensitive study for COD retrievals

4.3.1. Sensitive of properties at 415 nm

To build the lookup tables for COD retrievals using T, a sensitivity

analysis of T to changes in some variables was needed. Variations of T at 415

nm modeled by uvspec to variations of surface albedo, aerosol vertical profile,

SSA, g, AOD, gas concentrations and cloud properties are analyzed in this

section. The sensitivity of each variable was carried out by changing the

variable between percentiles 5 and 95 % according to measured data, while

keeping the others constant. Values of COD were assumed constant equal 2 for

ice clouds and 10 for low clouds. Default model configuration (section 3.4) is

used with CSZA of 0.86, AOD415 with 0.3, SSA415 assumed from SSA440 with

0.81 and g 415 assumed from g440 with 0.7.

Variations of T415 with respect to aerosol properties are presented in

Figure 18. The effect of the four aerosol profiles, retrieved from ground-based

measurement of Lidar and sun-photometer at the site, is compared with the two

profiles given by the LibRadtran library, labeled summer-spring and winter-

autumn (Mayer et al., 2015). It is clearly observed the little sensitivity of T to the

71

different aerosol profiles (Figure 18.a). But differences about 1.5 % were found

between the typical profile and the profile for spring.

Figure 18: Sensitivity of transmittance at 415 nm for liquid and ice clouds to variations of aerosol vertical profile [mo_ty= morning typical, mo_sp= morning in spring, af_ty=afternoon typical, af_sp=afternoon in spring, su-sp=summer-spring and au-wi=autumn-winter given by LibRadtran (Mayer et al., 2015) (a), variations of AOD415 (b), variations of SSA440 (c) and g440 (d).

In the case of variations of AOD415 (Figure 18.b) changes were observed,

especially for ice clouds, for which T415 decreased linearly with slope of about

21.5 % per AOD415 unit. For low clouds, there is less sensitivity to aerosol load

of about 8.3 % per AOD unit.

The SSA440 also influenced the T415 (Figure 18.c), the increase of SSA440

contributed to the increase of scattering and hence T415 was higher. Between

percentile 95 and 5, the differences of T were about 7 % for liquid and ice

clouds. The variation of T between 5 % and 95 % of g440 obtained from

AERONET products was almost imperceptible (Figure 18.d). Differences

between the extreme values of g440 were 0.6 % and 0.1 % for ice and low

clouds, respectively.

The sensitivity of T415 to CSZA was higher (Figure 19 a), especially for

ice clouds, between 95 % and 5 % percentiles, the difference can reach more

72

than 100 %. For surface albedo, NO2 and O3 there is no perceptible variation

(Figure 19).

Figure 19. As in the Figure 16 but for CSZA (a), surface albedo at 415 nm (b), dioxide of nitrogen (NO2) (c) and ozone (O3) (d).

Figure 20. Variations of total transmittance at 415 nm to cloud base, thickness, re and LWC for ice and liquid clouds with a constant value of COD.

Sensitivity of T415 to cloud properties is shown in Figure 20. For liquid

clouds, T415 varied around 2 % with the height of cloud base between

percentiles 5 and 95 %. Note the lower variability to re of ice and liquid clouds.

For high clouds, and re larger than 60 µm the shift, of about 1.3 % in T, is due to

73

the change of habit. For cloud thickness and IWC/LWC no sensitivity was

observed, as COD was assumed constant.

4.3.2. Errors affecting COD retrieval

From the previous analysis, the main variables to take into account for

retrieving COD are T415, AOD415, SSA, cloud properties such as cloud base

height (CB) and re as median values. In this section, the contribution of error of

each variable to COD retrieval is investigated.

In Figure 21, the relative error of each variable in the COD retrieval for low

and ice clouds is shown. For ice clouds errors are higher than low clouds.

Smaller errors below 1 % for low clouds properties are observed, but for ice

clouds the error of assuming a median of re can be 10 %

Note the high influence of aerosol optical properties for the error for COD

below 20. For COD below 5, errors are higher than 50 %. However for COD

above 50, the errors are lower than 10 % for low clouds. Comparing to the error

of the retrieval method of COD for low clouds using sun-photometry (17 %)

(Chiu et al., 2010), the current method has better accuracy considering total

overcast conditions for lower clouds, when COD is higher than 20.

Figure 21. Relative error of COD retrieval due to errors in each variable: AOD, SSA, T, re and CB and the total error for each COD value. For low clouds (a) and high clouds (b).

The current method does not include an estimation of instantaneous values

of spectral SSA. Therefore, for future works, consideration of SSA estimation in

the instantaneous measurements is advised. SSA can be retrieved using

74

measurements of spectral diffuse irradiance from the MFRSR for clear sky

conditions (Serrano et al., 2014) or can be obtained from sun-photometer

measurements (Dubovik et al., 2000).

4.3.2.1. Error of AOD estimation into COD retrieval

As pointed out previously, AOD415 is the second component that most

affects COD retrievals. Therefore, the validation of methods employed to infer

AOD415 in cloudy conditions and the contribution of the error to COD retrievals

are needed.

The accuracy of the method for inferring AOD is validated using

coincident AOD440 retrieved from sun-photometer. Note that methods do not

include the possible influenced of cloudiness on the AOD profile, as usually is

carried out in previous works (Guerrero-Rascado et al., 2013; Mateos et al.,

2013; Salgueiro et al., 2016), that is a topic where more studies will be needed

in the next years. For each method, some AOD440 cases in the sample are

assumed as not measured. These cases were estimated by the method and

differences with real AOD440 were calculated. Afterward, errors were computed

in reference to the real value.

The errors of each method are presented and resumed in Table 7. The

best accuracy was observed by time interpolation method on the same day

(error of 0.03). When AOD values on the day were not available, for instance a

value near sunrise or sunset, the estimation was performed using the nearest

value and adjusted using the mean monthly diurnal cycle. These estimations

resulted in errors between 0.07 and 0.08 (inter 3 and 4).

The worst accuracy was obtained using the monthly mean diurnal cycles

of AOD (error of 0.11). The method was applied only for cases when there were

no AOD measurements as close as two days, otherwise the monthly diurnal

cycle was adjusted using diurnal mean AOD of the nearest day. Note that the

accuracy improved using adjusted diurnal cycles, with errors of 0.08 (Inter 5).

The contribution of these errors to total error of COD is shown in Figure

22. From the figure, an improved of accuracy in the retrieval is observed

specially by the time interpolation throughout the day. Despite this improved in

75

estimating AOD for cloudy conditions, error of retrieval for lower CODs is

higher.

Table 7: Error of estimating AOD440 for cloudy conditions of each method: time interpolation throughout the day (Inter 1), monthly mean diurnal cycle of AOD (Inter 2), adjusted monthly mean diurnal cycle from the AOD measured in the previous time (Inter 3) and later time (Inter 4), adjustment of the climatological monthly diurnal cycle of AOD from mean diurnal AOD value computed as close as at least 2 days (Inter 5).

Method RMSE MAE Error R. Error Number

Inter 1 0.22 0.13 0.03 0.15 19979

Inter 2 0.60 0.46 0.12 0.62 20203

Inter 3 0.47 0.32 0.08 0.34 16034

Inter 4 0.41 0.28 0.07 0.32 16163

Inter 5 0.48 0.36 0.08 0.41 4328

Figure 22: Total relative error of COD retrieval, for different methods of assuming AOD in cloudy cases, for low clouds (a) and ice clouds (b).

4.3.3. Model response to clear sky global irradiance

As explained previously, pyranometer measurements were performed

every minute, while AOD retrievals from AERONET were available in intervals

of 15 minutes or longer. Thus, in this section, we compare measured irradiance

with the pyranometer with modeled by LibRadtran in clear sky conditions. Two

scenarios were analyzed: cases when there was temporal coincidence of

AOD500 and irradiance measurement (no interpolation) and cases when AOD500

was interpolated to pyranometer measurement time.

In Figure 23, the ability of the model to simulate global irradiance as

measured by the pyranometer for clear sky condition is shown initially

76

determined by all sky-camera. In order find out cases with presence of sub

visual cirrus in clear sky cases, from AOD level 1.0 to 1.5 cloud contaminated

cases are selected. Therefore the selected cases coincident in time with

pyranometer clear sky measurements were considered as cloud contaminated

and not taking into account in the comparison. A total of 14472 cases of clear

sky were analyzed.

The good agreement is observed between modeled and measured clear

sky G, with NRMSE of about 4 % and the overestimation of G of about 3 %.

Around 98 % of cases an absolute difference was below 10 % and 80 % of

them were lower than 5 % (Figure 22 b). Overestimation can be a consequence

of including a mean value of SSA, especially in spring with 0.9, because not

always in this season, biomass burning aerosols reach the region. Differences

were into the range of errors of pyranometer of about 5 %.

Figure 23: Instantaneous global irradiance modeled by 1-D uvspec and measured by pyranometer (a), frequency distribution of absolute differences between modeled and measured values (b), and relative differences using no interpolated AOD and water vapor (not interpolated) (c), interpolated on the day or adjusted from monthly diurnal cycle (interpolated_1) (d) and AOD and water vapor computed from mean monthly diurnal cycles (interpolated_2) (e).

Relative differences were computed for different AOD500 and water vapor

interpolation schemes and shown in Figures 23 c-e. Observe how there is no

significant differences between interpolation schemes. When no interpolation

was applied, the modeled result could reach a maximum of 12 % difference, but

more than 97 % of cases were below 10 %. Observe that only 1 % of cases

were underestimated by the model. Using interpolation on the same day, the

differences were similar to not interpolated cases, but there was a little increase

for CSZA below 0.2 with maximum of 15 %. It is noteworthy the increase in

77

underestimation of model for cases when interpolation was made using monthly

mean diurnal cycles, but differences remained between -5 % and 10 %.

Relative differences of Tdir and T at 415 nm of coincident modeled and

measured by reference MFRSR (numbered 345) are shown in Figure 24. A total

of 1105 clear sky cases observed by the sky camera were analyzed. The

difference was higher for lower CSZA with extremes observed near 0.4 with 18

% of relative differences for T and 12 % for Tdir.

Differences computed for direct sun transmittance allow having an

estimate of model error for computing direct sun transmittance at 415 nm. This

error will be employed for direct sun criterion for defining solar disk conditions.

The direct sun criterion assumed an error up to 15 %.

Figure 24: Relative differences of T at 415 nm modeled versus T at 415 nm measured

by MFRSR 345 (a,c) and of 𝑇𝑑𝑖𝑟 modeled and measured at 415 nm (b,d).

4.3.4. Spectral diffuse ratio validation.

Validation of DR is assessed in the following. In Figure 25 the DR for

cases of clear sky, low clouds with total overcast and high clouds with total

overcast conditions, carefully retrieved from camera observations, are shown. In

addition, the curve that limits clear sky conditions and cirrus clouds computed

by uvspec with AOD 0.3 is presented. Note the differences between cloudy and

78

clear sky conditions. Clear sky values were around the definition computed by

the model (green line), with lower ratio values, as the diffuse component under

clear sky is higher for shorter wavelengths, due to the influence of radiation

scattering by air molecules. When clouds are present, the component of

Rayleigh scattering becomes less important, with an increase of multiple

scattering due to clouds and Mie scattering, decreasing the spectral

dependence.

Figure 25: Validation of ratio of downwelling diffuse irradiance (D↓) at 870 nm and 415 nm (DR) for cases discriminated by the all-sky camera (a), the classification defined from model and observed values (c), for a specific day with low broken clouds (b) and for a specific day of low clouds with total overcast condition (d).

High clouds showed intermediate values, i. e., above clear sky threshold

and below low cloud values, but some overlaps can be observed. It is important

to note that only visual cirrus clouds are shown, leading to higher differences

with clear sky. It is remarkable how low cloud ratios are systematically above

0.58 and no sensitive to CSZA. In the case of low clouds, besides the little

spectral dependence, the contribution of surface albedo favors to increase the

differences with clear sky. Under these considerations, it is possible to classify

as clear sky or cloudy atmosphere. In the figure 25 c definitions are shown

based on the former figure and Figure 5.

79

Specific days of study with low cloud overcast conditions and low broken

clouds are shown in Figures 25 b-d. Observe the differences between the

cases, for low broken clouds the spectral ratio never reached 0.58 compared to

total overcast where about 99 % of the cases are above 0.58. That can be a

good indicator of cloud cover as reported by Min et al. (2008). Then, based on

the analyses just shown, it is possible to use the spectral ratio of diffuse

irradiance, at 870 nm and 415 nm, to define conditions of cloudiness. These

conditions were used especially for observations when the sky camera was not

available, i.e. before September of 2016.

4.4. Effective cloud optical depth (ECOD)

4.4.1. Validation with pyranometer ground-based

measurements

A total of 6870 cases of ECOD retrieved coincident in time with

pyranometer measurements were used as input for computations of G. The

irradiance calculated by the uvspec model was compared with pyranometer

measurements. For this analysis, ECOD values were modeled with different

effective radius of values from percentiles 5 and 95% of re retrieved from

CloudSat-CALIPSO products. For low clouds the relation between modeled and

measured is shown in Figure 26.

Lower differences are observed with the increase of CSZA, as expected.

For CSZA near 0.2 higher differences up to 100 % are observed. Differences

with maximum of 10 % of absolute value are observed for CSZA higher than

0.95. A good agreement between modeled and measured results is observed

with NRMSE of 9 % for the three re used; more than 60 % of cases are below

10 % of absolute difference. But using percentile 5 of re, there is a slightly

underestimation of the model, and for the 95 % percentile, little overestimation

is observed with MBD equal 0.03.

80

Figure 26: Comparison of modeled global irradiances using percentile 50 (a,d,g), percentile 5 (b,e,h) and percentile 95 of re for ECOD values retrieved for low clouds with total overcast condition.

For cirriform clouds (Figure 27), 744 cases were analyzed. Despite the

error retrieving ECOD for high clouds is higher than low clouds, differences

observed for high clouds are lower as compared to low clouds. That can be due

to the fact that, for high clouds, D↓ has less weight in G because SS is not

totaling extinct and 1-D model computes SS with higher accuracy as compared

to D↓. No significant difference is observed with re variation, with the best

agreement for data corresponding to percentile of 95 %.

4.4.2. Comparison with COD retrieved from sun-photometer

The sun-photometer (AERONET) method to estimate COD is based on

zenith radiance measurements and not from hemispheric measurements as for

MFRSR. The AERONET method is more sensitive to the low scale variation of

cloudiness (Chiu et al., 2010) and even narrow interstices at zenith can

contribute to differences. The method is more accurate in the presence of low

broken clouds, due to the higher dependence on surface albedo, but for low

81

clouds, in overcast conditions, the contribution of surface albedo can be

reduced, resulting in larger differences.

Figure 27: As Figure 25 but for cirrus clouds with total overcast.

In Figure 28, a comparison of COD retrieved from MFRSR every 1 min

and AERONET, with each measurement taking about 1.5 min, every 15

minutes, at the site, is shown. For a specific day, with low clouds and total

overcast conditions, coincident retrievals are presented (Figure 27 a). Note the

CODs agreement between the two datasets and the variability observed for 1

min resolution from MFRSR with COD peaks every 10 or more minutes.

A total of 190 cases of low clouds with total overcast conditions were

carefully selected from the sky camera and conditions of obstruction of direct

solar radiation. For this 190 time coincident COD cases, shown in Figure 28 b,

differences were observed with mean averaged error of 23 % and

underestimations of MFRSR COD of 18 %. The discrepancy was higher for

COD values above 40, where underestimation of MFRSR respect to AERONET

is clearly observed. Those differences have to do with the different procedures

82

of each method. COD from AERONET is more sensitive to local variations of

cloud properties while ECOD from MFRSR depends on cloud cover of the sky.

Figure 28: Comparison between COD retrieved from AERONET and the COD retrieved by MFRSR under the presence of low clouds in total overcast condition. Time series for one specific day (a), COD of AERONET vs COD of MFRSR for coincident cases (b). G modeled using COD AERONET vs G measured by pyranometer (c). G modeled using COD from MFRSR vs G measured by pyranometer (d). Differences with respect to CSZA for AERONET (e) and MFSRSR (f). Frequency distribution of absolute error for AERONET (g) and MFRSR (h).

The coincident CODs were included as input in the uvspec model and

solar global irradiances at the surface were calculated and compared with

values measured by the pyranometer (G).

Better agreement between modeled and measured G was observed when using

COD estimated from MFRSR (this work) (Figure 28 d). For MFRSR, the

differences were lower for higher values of CSZA. For AERONET, no

dependence with CSZA was observed. Differences under 10 % for MFRSR

represented about 70 % of cases, however only 32 % for AERONET. In

83

resume, it was observed that COD retrieved by MFRSR under total overcast

conditions allowed computations of G at surface that agreed better with

measured G than AERONET COD in the same conditions.

4.4.3. Effective cloud optical depth, seasonal and diurnal

variability.

The seasonal and diurnal COD distribution for low clouds in total overcast

condition is shown in Figure 29. Furthermore, histogram of frequency, some

statistical values and number of cases are also shown. Statistic differences

between season and hours data are observed and checked with the Kruskal-

Wallis hypothesis-test (Kruskal and Wallis, 1952) .

Total overcast scenarios due to low clouds were more (less) frequent in

spring (autumn). In winter (Figure 29.c), lower values of COD were observed,

with the lowest median of 21.2, with 20 % of the distribution with COD values

above 40. In spring, there was an increase of COD above 40, with 32 % of the

cases. The median value for this season was 28.6 being the maximum of all

seasons. Between summer and autumn there was no remarkable differences,

with median around 24, but, for autumn, percentile 95 was the largest, equal to

86.5.

Also, the diurnal variability of COD was observed (Figure 29.e-h). Near

sunrise, values of COD were smaller with the lowest median of 19.0. The

median value increased until the time interval between 12:00 and 14:00 LH

(28.3) with 30 % of data with COD above 40. Near sunset the median

decreased to 25.9, but still almost 30% of the data presented COD above 40.

The analysis was also applied for visible cirrus cloud (Figure 30) COD.

The number of cases decreased as compared with low clouds. As for low

clouds, the differences were observed for diurnal and seasonal time intervals,

and checked the statistical significance by The Kruskal-Wallis hypothesis test.

Cirrus clouds presented COD values above 5, especially in summer, with

23.6 % of the cases. Higher values were observed in spring with median of 3.11

followed by summer with median of 2.88. In autumn, the lowest median was

observed and with only 2.5 % of the cases above 5, for the 400 available cases.

Lower values of COD were also observed for sunrise with median of 1.94.

84

Maximum median occurred near midday with 3.21 but after 14:00 LT, the

highest amount of cases of COD above 5 was observed (19.6 % of cases).

Figure 29. Seasonal and hourly COD distribution for low clouds. For summer (a), autumn (b), winter(c), spring (d). From sunrise up to 08:00 LT (e), between 08:00 and 10:00 LT (f), around noon (from 10:00 up to 14:00 LT) (g) and after 14:00 LT(h).

Figure 30. As Figure 28 but for cirrus clouds.

85

4.5. Cloud effects on solar radiation

4.5.1. Shortwave cloud radiative effects

The cloud effects on solar radiation (CRE and NCRE) are shown in the

figure 31 for different cloud scenarios defined in the table 3. Considering that

the model generally overestimated the global irradiance, enhancement effects

were underestimated. The criterion applied to define enhancement was based

on a threshold computed from relative error of the model to compute clear sky

G (section 3.5.4). Also, cloud effects were classified depending on the solar

disk, i.e. if it was clear or blocked by clouds (defined by 1 and 0, respectively).

When the solar disk was clear, enhancement effects were observed for

the three cloud types analyzed, with the highest effect for low clouds with

median of CRE (NCRE) of 79 W/m2 (0.10). High clouds had the least

enhancement effect with median of CRE (NCRE) of 49.8 W/m2 (0.08). It is

important to note that enhancement effect for mid clouds was comparable to

lower clouds with CRE (NCRE) of 75.2 W/m2 (0.12), but only 32 cases were

found.

Extreme enhancement cases for CRE (NCRE) were found for low clouds

reaching near 200 W/m2 (0.3). The extremes of CRE were produced in isolated

cases with 75 % observed between January and March with maximum in March

with 30 % of cases. They were most prevalent (77 % of cases) between CSZA

values of 0.85 and 1.

NCRE was less dependent on CSZA and NCRE cases above 0.3 were

found throughout the year, with the higher number of cases in May (35 % of

total of extreme cases). Near CSZA 0.3, the majority of cases were observed,

with 28 % of total of cases. As Marín et al. (2017) reported, the enhancement

effect does not depend on CSZA and the maximum of cases were observed

between 40-50 % of cloud cover. They reported a median of NCRE of 0.18 and

percentile 75 of 0.3.

An opposite cloud effect (shortwave radiation attenuation) was observed

when clouds obstructed the solar disk. The maximum deficit was found for low

86

clouds in total overcast condition, with median for CRE (NCRE) of -400 W/m2 (-

0.72) and maximum of -880 W/m2 (-0.99). For low clouds with no total overcast

condition (LWBK_0), a less negative effect was found for NCRE, with median of

-0.68. For low clouds, specially cases of broken clouds, an increase of diffuse

irradiance is expected due to 3D effects related to conditions of low broken

clouds (Tzoumanikas et al., 2016).

Figure 31: Cloud effects computed for cloudiness conditions and solar disk. Subscript 0 indicate obstructed solar disk and 1 not obstructed by clouds. Cloudiness are LW=low clouds, LWBK=low broken clouds, H=High clouds with total overcast, HBK=High broken clouds, MID=mid-level clouds with total overcast, MID_BK =mid-level broken clouds.

Mid-level clouds with total overcast had less attenuation effects

compared to low clouds but the maximum could reach similar values. The

median for mid clouds, for CRE and NCRE, were -240 W/m2 and -0.57,

respectively. High clouds, as expected, had the lowest cooling effects with

median -190 W/m2 and -0.33 for CRE and NCRE, respectively. Note, extremes

of cooling effects of high clouds with NCRE (CRE) around -0.7 (-450 W/m2)

comparable to median of low clouds (LW_0).

Four cases of days under low broken clouds (Figure 32.b, d), total

overcast of low clouds (Figure 32 c) and clear sky (Figure 32 a) conditions were

selected from visual inspection of sky camera pictures. G measured by

pyranometer, G modeled under clear sky, upper threshold of G modeled (in

Figure 32 a,b,d) and lower threshold of G modeled (in Figure 32 c) due to error

of model (section 4.3.3) and broadband SS from MFRSR are shown. SS from

87

MFRSR was employed to visually discriminate when the sun was totally blocked

or not by clouds.

For clear sky conditions, SSA440 mean value of 0.85 retrieved on the

same day by sun-photometer was employed. The good response of the model

and the systematic overestimation relative to measurements can be observed.

The mean relative error observed is around 5 % with a maximum of 12 %

observed for the lower G values near 16:00 LT.

Due to this overestimation, the instants with enhancement effects might

be slightly underestimated. For the case of total overcast condition (Figure

32.c), G measured by the pyranometer is clearly below the lower limit of the

modeled G for clear sky condition minus one standard deviation.

The behavior observed for low broken clouds in Figure 32.b and Figure

30 d shows peaks of enhancement of G lasting, at maximum, about 20 minutes

(Figure 32.d between 15:30-15:50). These cases of enhancement were

observed when direct-sun was not blocked and they occurred at any time along

the day.

Figure 32: Specific day with clear sky conditions (a), low broken clouds (b,d) and total overcast (c).

88

4.5.2. Cloud efficiency

The dependence of the shortwave cloud radiative effect variables, CRE

and NCRE, on COD and CSZA is presented in Figure 32 for liquid and ice

clouds. The strong dependence of CRE on CSZA is observed, with lower CRE

(higher negative effect or attenuation) for higher CSZA values.

The coefficient values (m =linear slope and R2) of linear fit between CRE

/NCRE and ln (COD) for CSZA centered at 0.32 to 0.97 with a range of ±0.025

were computed. One observes the good fit of CRE/NCRE with ln (COD), being

less scattered as CSZA increases, especially for liquid clouds (Mateos et al.,

2014). In addition, the slope became more negative for higher CSZA. A better

accuracy was found when NCRE was used instead of CRE for liquid clouds.

Furthermore, the NCRE vs ln (COD) was less dependent on CSZA (m less

variable to changes on CSZA).

For ice clouds, the linear fit could be applied for lower COD (below 3), but

with the increase of COD the relationship with ln (COD) was better observed.

Therefore, to avoid discrepancies between ice and liquid clouds the same linear

fit was applied, even though the relationship was not as robust as for liquid

clouds. In addition, using NCRE for ice clouds the dependence with CSZA was

not minimized.

The linear fit computed for liquid clouds was not as good as obtained by

Mateos et al. (2014), because they computed COD from pyranometer

measurements with less accuracy that the current work. But the relationship

was still strong in this work, which allowed computing the radiative cloud

efficiency from CRE. On the other hand, cloud efficiency could be computed

from NCRE (Salgueiro et al., 2016) for liquid clouds with better accuracy. Due to

the lower precision observed for ice clouds, CEF will be computed only for

CSZA equal 0.85, in order to make a comparison with liquid clouds. CEF was

computed using the slope of the linear fit for each CSZA interval (equation 30)

and the correspondent value of COD. CEF computed from CRE (Figure 33 a)

presented sensitivity to CSZA and COD. For any value of COD, the lower CSZA

the lower CEF (in absolute units). Note the strong dependence on COD. For

lower COD direct sun irradiance was not completely extinguished, and the

89

radiative effect included direct sun irradiance at the surface. But with the

increase of COD, direct sun irradiance became less important until it was totally

extinct (around COD 20). For higher COD, only the diffuse irradiance was

observed at the surface. Because of that, the lower decrease of CEF for COD

above 40 is observed.

Figure 33. Relationship between CRE vs COD for liquid clouds (a) and ice clouds (c).

Relationship between NCRE and COD for liquid clouds (b) and ice clouds (d). Colors are related to central values of CSZA. Value of m is the slope and R

2 the determination coefficient

of the curve Y =m lnCOD + n.

Maximum of CEF computed from CRE was of -65 Wm-2 per COD-unit for

values of CSZA near 0.35 and 0 COD, with the minimum CEF of almost 0 Wm-2

COD-unit-1 for the higher COD. For liquid clouds, the CEF computed from

NCRE only depended on COD, the notable decrease observed until COD 40 is

observed. After 40, it is almost invariable, with CEF near -0.01 COD unit-1.

Differences can be observed between liquid and ice clouds for CEF

computed from NCRE (Figure 33 c) in the COD interval between 0 and 12, for

90

CSZA around 0.85. Low clouds showed less efficiency as compared to ice

clouds, with the maximum difference of 30 % around COD equal 2. Low clouds

have higher values of asymmetry factor than ice clouds with 3D ice particle type

(one example is Rough-aggregate used in the uvspec model) in the spectral

range from 280 nm up to 1500 nm (see Figure 2). Thus, the forward scattering

peak is stronger as compared to ice clouds.

Figure 34. Cloud efficiency computed for CRE (a) and NCRE (b) for different CSZA

values as function of COD for liquid clouds. Comparison between CEF (from NCRE) vs COD for liquid and ice clouds, for CSZA centered at 0.85 (c). Seasonal variation of CEF (from NCRE) for liquid clouds (d).

In addition, comparing the parametrization employed for retrieving COD

for ice and low clouds around 415 nm, because there is no absorption of

radiation by clouds (Figure 2), for the same T, lower COD is observed for ice

clouds compared to low clouds. Moreover, ice clouds have higher absorption

91

efficiency in the spectral range between 1000 nm and 2000 nm, thus surface

albedo can have a stronger impact on low clouds.

Those processes aforementioned can contribute to the lower efficiency of

low clouds over ice clouds. There is no evidence of variation of CEF with

respect to the season (Figure 33), but more detailed analysis is needed, for

example the screening of cases with low aerosol loads versus polluted cases.

92

5. Conclusions

Variability of cloud profile at the region of MASP was observed by using a

combined product from CALIPSO-CloudSat. Higher frequency was found in

summer when clouds could reach 17 km, but, in winter, clouds could reach only

15 km. Differences between daytime and nighttime profiles were also detected.

At daytime profile, the maximum frequency of around 30 % took place near

2500 m in summer. At night, only in winter the cloud frequency did not develop

at higher layers. In addition, the increase of the height of clouds inland, as

compared to the coast, was notable. Near the coast, clouds frequency had

maximum around 2 km, and inland, as far as 100 km from the coast, the

maxima were observed near 11 km altitude. The predominance of ice-3D cloud

particle was found, especially for clouds above 10 km.

Cloud climatology for São Paulo, Brazil, for the period from 1958 to 2016

using hourly visual observations was performed. The diurnal cycle of cloud

amount (cloud cover) was dominated by low clouds especially stratiform low

clouds, with maximum at sunset and decreasing observed after sunrise until

about 11:00 local standard time. Almost all cloud types showed an inverse

diurnal cycle compared to the diurnal cycle of the solar heating, except low

cumuliform clouds. Remarkable differences between cumuliform and stratiform

clouds were also observed. The maximum presence of Sc in the afternoon,

particularly in summer, can be related to complex dynamical mechanisms acting

in the region. Its location near the coast where the sea breeze can interact with

the heat island effect is an example, but further studies are necessary to a

complete understanding of Sc formation and development in the region.

The diurnal cycle of clouds varied along the year. Larger differences

were observed between April-August and the rest of the year. Between April

and August the variation of the diurnal cycle amplitude was associated to the

variation observed for stratiform clouds, whereas for summer-spring, the

variation of the amplitude of cumuliform clouds, mid-level and high clouds was

higher. The maximum cloud coverage was observed in the afternoon for

summer-spring, but at sunrise in April-August for all cloud types. The annual

93

cycle showed the decrease of cloud amount of all cloud types in winter, our dry

season, but maxima were variable and could be observed in spring or summer.

Seasonal and annual trends of cloud amount were computed considering

the first 30 years, the last 30 and for the entire period (58 years). Significant

trend for total cloud cover was only observed in autumn for the whole period

(0.8 %/decade) and in the first 30 years (2.3 %/decade). Annual mean low cloud

amount increased by 1.6 %/decade, due to the increase of 3.1 %/decade for

stratiform low clouds, particularly in the first 30 years. Annual mean mid-level

cloud amount decreased with a trend of -2.4 %/ decade. For the case of

cirriform clouds, a statistically significant positive trend was observed only in

autumn. In the annual basis, the trend was about 0.8 %/decade. For the entire

period, the decrease of total cloud cover at sunrise (-0.5 %/ decade), and an

increase of 0.8 %/ decade at midday were observed. Low cloud cover increased

after midday (2.2 %/ decade), while cirriform clouds increased (0.8 %/ decade)

only around midday. Mid-level clouds decreased in all the analyzed periods of

the day.

Comparing with the global climatology, there was an agreement in trend

for mid and high level clouds for the period of 1971-2009, but for low clouds,

significant differences were observed, probably due to local effects influencing

the characteristics of low stratiform clouds.

For retrieving COD and cloud effects for total overcast conditions, cloud

scenarios were defined by the coincidence of three criteria: visual observation,

the ratio of spectral diffuse irradiance between 870 nm and 415 nm channels

and direct solar radiation blocked by clouds. Moreover, from September 2016,

visual observations using the sky-camera were included.

The effective cloud optical depth (ECOD) was retrieved at the site using

the total hemispheric transmittance at 415 nm estimated from the MFRSR. The

method presented good accuracy for low clouds, with error below 15 % for COD

above 20. But for ice clouds the error was higher, around 20 % for COD equal

20. For COD below 2, for both cloud types, errors were higher than 80 %

associated to uncertainties in the optical properties of aerosols. Therefore the

method was more efficient for low clouds. Consequently, methods for retrieving

94

COD with better estimation of aerosol single scattering albedo and using direct-

sun measurements for lower COD can contribute to decreasing the error.

The values of ECOD presented seasonal variability, with the lowest for

low clouds in winter with median of 21.2 and maximum in spring, equal 28.6.

For cirrus clouds, the minimum was observed in autumn with median of 2.21

and maximum in spring with 3.11. In addition, for low clouds (cirrus) near

sunrise the ECOD values were lower with median of 18.9 (1.94) as compared to

the afternoon, with 28.3 (3.21). Values of ECOD for cirrus clouds could reach 10

in few cases.

Cloud effects on solar radiation at the surface were computed by the

normalized shortwave cloud radiative effect (NCRE) and cloud radiative effect

(CRE). Using the NCRE, the surface albedo was not taking into account and it

was less sensitive to CSZA. Solar global irradiance at the surface under cloudy

conditions was measured by a pyranometer and clear sky irradiance was

computed by the RTM. Cloud radiative effects depended on the conditions of

the solar disk, cloud type and cloud cover. When the solar disk was blocked by

clouds, low cloud and total overcast conditions, NCRE presented a median

value of -0.72, but when the sky was not totally overcast, the median was -0.68.

For mid and high clouds the median of NCRE was -0.57 and -0.33, respectively.

The enhancement effect was observed when the solar disk was clear for low

(0.1), mid (0.12) and high clouds (0.07), with extreme values for low clouds

(above 0.4). This enhancement effect could last up to 20 minutes.

The shortwave cloud efficiency (CEF) of low clouds was maximum for

lower CSZA and COD. With higher values of COD no variation of CEF was

observed with values near 0. The efficiency of low clouds was lower as

compared to ice clouds. Differences up to 30 % for values of COD near 2 were

estimated.

95

6. Future works

Application of the method to retrieve COD of semitransparent clouds

using direct sun transmittance (Min, 2004) is needed. The method could

improve the accuracy of COD especially for lower values, and can

contribute to better understand the radiative characteristics of cirrus

clouds at the region.

Effective radius (re) is one of the most important microphysical properties

of clouds employed in the radiative transfer models. Variation of this

parameter can lead to important variations of the impact of clouds in the

radiation budget. Satellite derived products of re have limitations

especially for low clouds. Specifically, under polluted conditions this

variable can be affected particularly for low clouds. Therefore, continuous

retrievals of this variable for the region are important.

The use of 1-D RTM has some limitations when evaluating the cloud

effects on the radiation budget e. g. cloud properties can only be

retrieved under total overcast conditions or simple relationships with sky

fraction for computing COD for partially overcast conditions. That is far

from the actual configuration of cloud fields usually found. Therefore

employing 3-D RTM in order to evaluate the 3D effects and, in synergy

with sky camera, optical properties of clouds can be retrieved in partially

overcast conditions, even when the sun disk is not blocked. Since

computations using 3D RTM can be very costly, the development of

parametrization of 3D for use in 1D-code can be an important step.

Due to the subjective nature of visual observations of cloud cover, it

would be important to test other long-term datasets of cloud cover from

ground-based or satellites measurements with enough spatial resolution.

Such measurements can be employed for comparison with results

obtained from visual observations. One example is the method

developed by Min et al. (2008), using ground-based measurements of

MFRSR at spectral wavelengths centered at 870 nm and 415 nm.

96

7. References

Ackerman, S. A., and Stephens, G. L. (1987). The Absorption of Solar Radiation by Cloud Droplets: An Application of Anomalous Diffraction Theory. Journal of the Atmospheric Sciences, 44(12), 1574–1588. https://doi.org/10.1175/1520-0469(1987)044<1574:TAOSRB>2.0.CO;2

Albrecht, B. A. (1989). Aerosols, Cloud Microphysics, and Fractional Cloudiness. Science, 245(4923), 1227. https://doi.org/10.1126/science.245.4923.1227

Alizadeh-Choobari, O., and Gharaylou, M. (2017). Aerosol impacts on radiative and microphysical properties of clouds and precipitation formation. Atmospheric Research, 185, 53–64. https://doi.org/10.1016/j.atmosres.2016.10.021

Almeida, M. P., Zilles, R., and Lorenzo, E. (2014). Extreme overirradiance events in São Paulo, Brazil. Solar Energy, 110, 168–173. https://doi.org/10.1016/j.solener.2014.09.012

Anderson, G. P., Clough, S. A., Kneizys, F. X., Chetwynd, J. H., Shettle, E. P., U.S. Air Force Geophysics Laboratory., and Optical Physics Division. (1986). AFGL atmospheric constituent profiles (0-120km).

Andrejczuk, M., Gadian, A., and Blyth, A. (2014). Numerical simulations of stratocumulus cloud response to aerosol perturbation. Atmospheric Research, 140–141, 76–84. https://doi.org/10.1016/j.atmosres.2014.01.006

Austin, R. T. (2007). Level 2B radar-only cloud water content (2B-CWC-RO) process description document. CloudSat Project Report. Retrieved from ftp://ftp.cira.colostate.edu/ftp/CloudSat/Docs/2B-CWC-RO_PDICD.P_R04.20071021.pdf

Baldridge, A. M., Hook, S. J., Grove, C. I., and Rivera, G. (2009). The ASTER spectral library version 2.0. Remote Sensing of Environment, 113(4), 711–715. https://doi.org/10.1016/j.rse.2008.11.007

Barnard, J. C., and Long, C. N. (2004). A simple empirical equation to calculate cloud optical thickness using shortwave broadband measurements. Journal of Applied Meteorology, 43(7), 1057–1066.

Boersma, K. F., Eskes, H. J., and Brinksma, E. J. (2004). Error analysis for tropospheric NO 2 retrieval from space: ERROR ANALYSIS FOR TROPOSPHERIC NO 2. Journal of Geophysical Research: Atmospheres, 109(D4), n/a-n/a. https://doi.org/10.1029/2003JD003962

Boucher, O., Randall, D., Artaxo, P., Bretherton, C., Feingold, G., Forster, P., Kerminen, V.-M., Kondo, Y., Liao, H., and Lohmann, U. (2013). Clouds and aerosols. In Climate change 2013: the physical science basis. Contribution of Working Group I to the Fifth Assessment Report of the Intergovernmental Panel on Climate Change (pp. 571–657). Cambridge University Press.

Brückner, M., Pospichal, B., Macke, A., and Wendisch, M. (2014). A new multispectral cloud retrieval method for ship-based solar transmissivity measurements. Journal of Geophysical Research: Atmospheres, 119(19), 11,338-11,354. https://doi.org/10.1002/2014JD021775

97

Burleyson, C. D., Long, C. N., and Comstock, J. M. (2015). Quantifying Diurnal Cloud Radiative Effects by Cloud Type in the Tropical Western Pacific. Journal of Applied Meteorology and Climatology, 54(6), 1297–1312. https://doi.org/10.1175/JAMC-D-14-0288.1

Carvalho, L. M., Jones, C., and Liebmann, B. (2004). The South Atlantic convergence zone: Intensity, form, persistence, and relationships with intraseasonal to interannual activity and extreme rainfall. Journal of Climate, 17(1), 88–108.

Chiu, J. C., Huang, C.-H., Marshak, A., Slutsker, I., Giles, D. M., Holben, B. N., Knyazikhin, Y., and Wiscombe, W. J. (2010). Cloud optical depth retrievals from the Aerosol Robotic Network (AERONET) cloud mode observations. Journal of Geophysical Research, 115(D14). https://doi.org/10.1029/2009JD013121

Christensen, M. W., Stephens, G. L., and Lebsock, M. D. (2013). Exposing biases in retrieved low cloud properties from CloudSat: A guide for evaluating observations and climate data. Journal of Geophysical Research: Atmospheres, 118(21), 12,120-12,131. https://doi.org/10.1002/2013JD020224

da Silva, J. J., Lopes, F. J. S., and Landulfo, E. (2017). Cloud-Aerosols interactions by multiple scenarios approach. In A. Comerón, E. I. Kassianov, and K. Schäfer (Eds.) (p. 33). SPIE. https://doi.org/10.1117/12.2278588

de Almeida Castanho, A. D., Vanderlei Martins, J., and Artaxo, P. (2008). MODIS Aerosol Optical Depth Retrievals with high spatial resolution over an Urban Area using the Critical Reflectance. Journal of Geophysical Research: Atmospheres, 113(D2), D02201. https://doi.org/10.1029/2007JD008751

de Miranda, R. M., Lopes, F., do Rosário, N. É., Yamasoe, M. A., Landulfo, E., and de Fatima Andrade, M. (2017). The relationship between aerosol particles chemical composition and optical properties to identify the biomass burning contribution to fine particles concentration: a case study for São Paulo city, Brazil. Environmental Monitoring and Assessment, 189(1). https://doi.org/10.1007/s10661-016-5659-7

Dubovik, O., and King, M. D. (2000). A flexible inversion algorithm for retrieval of aerosol optical properties from Sun and sky radiance measurements. Journal of Geophysical Research: Atmospheres, 105(D16), 20673–20696.

Dubovik, O., Smirnov A., Holben B. N., King M. D., Kaufman Y. J., Eck T. F., and Slutsker I. (2000). Accuracy assessments of aerosol optical properties retrieved from Aerosol Robotic Network (AERONET) Sun and sky radiance measurements. Journal of Geophysical Research: Atmospheres, 105(D8), 9791–9806. https://doi.org/10.1029/2000JD900040

Eastman, R., and Warren, S. G. (2013). A 39-Yr Survey of Cloud Changes from Land Stations Worldwide 1971–2009: Long-Term Trends, Relation to Aerosols, and Expansion of the Tropical Belt. Journal of Climate, 26(4), 1286–1303. https://doi.org/10.1175/JCLI-D-12-00280.1

Eastman, R., and Warren, S. G. (2014). Diurnal Cycles of Cumulus, Cumulonimbus, Stratus, Stratocumulus, and Fog from Surface

98

Observations over Land and Ocean. Journal of Climate, 27(6), 2386–2404. https://doi.org/10.1175/JCLI-D-13-00352.1

Freitas, E. D. (2003). Circulações Locales em São Paulo e sua influencia sobre disperção de poluentes (Doctoral Thesis). University of São Paulo,Brazil.

Freitas, ED, Silva Dias, P., Carvalho, V., Rocha, C., Martins, L., Martins, J., and Andrade, M. (2009). Factors involved in the formation and development of severe weather conditions over the megacity of São Paulo. In 8th Symposium on the Urban Environment, Modeling and Forecasting in Urban Areas. Proceedings of the 89th AMS Annual Meeting.

Gasteiger, J., Emde, C., Mayer, B., Buras, R., Buehler, S. A., and Lemke, O. (2014). Representative wavelengths absorption parameterization applied to satellite channels and spectral bands. Journal of Quantitative Spectroscopy and Radiative Transfer, 148, 99–115. https://doi.org/10.1016/j.jqsrt.2014.06.024

Guerrero-Rascado, J. L., Costa, M. J., Silva, A. M., and Olmo, F. J. (2013). Retrieval and variability analysis of optically thin cloud optical depths from a Cimel sun-photometer. Atmospheric Research, 127, 210–220. https://doi.org/10.1016/j.atmosres.2012.10.025

Gueymard, C. A. (2004). The sun’s total and spectral irradiance for solar energy applications and solar radiation models. Solar Energy, 76(4), 423–453. https://doi.org/10.1016/j.solener.2003.08.039

Hagihara, Y., Okamoto, H., and Yoshida, R. (2010). Development of a combined CloudSat-CALIPSO cloud mask to show global cloud distribution. Journal of Geophysical Research: Atmospheres, 115(D4), D00H33. https://doi.org/10.1029/2009JD012344

Hamed, K. H., and Rao, A. R. (1998). A modified Mann-Kendall trend test for autocorrelated data. Journal of Hydrology, 204(1–4), 182–196.

Heymsfield, A. J., Krämer, M., Luebke, A., Brown, P., Cziczo, D. J., Franklin, C., Lawson, P., Lohmann, U., McFarquhar, G., Ulanowski, Z., and Van Tricht, K. (2017). Cirrus Clouds. Meteorological Monographs, 58, 2.1-2.26. https://doi.org/10.1175/AMSMONOGRAPHS-D-16-0010.1

Holben, B. N., Eck, T. F., Slutsker, T. I., Tanre´, D., Buis, J. P., Setzer, K. A., Vermote, E., Reagan, J. A., Kaufman, Y. J., Nakajima, T., … Smirnov, A. (1998). AERONET—A Federated Instrument Network and Data Archive for Aerosol Characterization.

Holton, J. R., and Curry, J. A. (Eds.). (2003). Encyclopedia of Atmospheric Sciences (Edição: 1). Amsterdam ; Boston: Academic Press.

Houze, R. A., JR. (2014). Cloud Dynamics: 104 (Edição: 2). Amsterdam ; New York: Academic Press.

Hu, Y. X., and Stamnes, K. (1993). An Accurate Parameterization of the Radiative Properties of Water Clouds Suitable for Use in Climate Models. Journal of Climate, 6(4), 728–742. https://doi.org/10.1175/1520-0442(1993)006<0728:AAPOTR>2.0.CO;2

Hui, W. J., Cook, B. I., Ravi, S., Fuentes, J. D., and D’Odorico, P. (2008). Dust-rainfall feedbacks in the West African Sahel: DUST-RAINFALL FEEDBACKS. Water Resources Research, 44(5). https://doi.org/10.1029/2008WR006885

Joiner, J., Vasilkov, A. P., Bhartia, P. K., Wind, G., Platnick, S., and Menzel, W. P. (2010). Detection of multi-layer and vertically-extended clouds using

99

A-train sensors. Retrieved from http://ntrs.nasa.gov/search.jsp?R=20110014293

Key, J. R., Yang, P., Baum, B. A., and Nasiri, S. L. (2002). Parameterization of shortwave ice cloud optical properties for various particle habits. Journal of Geophysical Research: Atmospheres, 107(D13), AAC 7-1. https://doi.org/10.1029/2001JD000742

Kruskal, W. H., and Wallis, W. A. (1952). Use of Ranks in One-Criterion Variance Analysis. Journal of the American Statistical Association, 47(260), 583–621. https://doi.org/10.2307/2280779

Kurucz, R. L. (1994). Synthetic Infrared Spectra (pp. 523–531). Dordrecht: Springer Netherlands.

Landulfo, E., Larroza, E. G., Lopes, F. J. S., and Hoareau, C. (2009). A preliminary classification of cirrus clouds over Sao Paulo city by systematic lidar observations and comparison with CALIPSO and AERONET data. In R. H. Picard, K. Schäfer, A. Comeron, and M. van Weele (Eds.) (p. 747506). https://doi.org/10.1117/12.829923

LeBlanc, S. E., Pilewskie, P., Schmidt, K. S., and Coddington, O. (2015). A spectral method for discriminating thermodynamic phase and retrieving cloud optical thickness and effective radius using transmitted solar radiance spectra. Atmospheric Measurement Techniques, 8(3), 1361–1383. https://doi.org/10.5194/amt-8-1361-2015

Lee, H. (2016). Analysis of Carbon Dioxide and Cloud Effects on Temperature in Northeast China. Procedia Computer Science, 91, 690–699. https://doi.org/10.1016/j.procs.2016.07.172

Leontyeva, E., and Stamnes, K. (1994). Estimations of Cloud Optical Thickness from Ground-Based Measurements of Incoming Solar Radiation in the Arctic. Journal of Climate - J CLIMATE, 7(4), 566–578. https://doi.org/10.1175/1520-0442(1994)007&lt;0566:EOCOTF&gt;2.0.CO;2

Liou, K. N. (2002). An Introduction to Atmospheric Radiation (Edición: 2). Amsterdam ; Boston: Academic Press.

Mace, G. (2007). Level 2 GEOPROF product process description and interface control document algorithm version 5.3. NASA Jet Propulsion Laboratory.

Mahoney, W. P. (1988). Gust Front Characteristics and the Kinematics Associated with Interacting Thunderstorm Outflows. Monthly Weather Review, 116(7), 1474–1492. https://doi.org/10.1175/1520-0493(1988)116<1474:GFCATK>2.0.CO;2

Manton, M. J. (1980). Computations of the effect of cloud properties on solar radiation. Journal de Recherches Atmosphériques, 14, 1–16.

Marchand, R., Mace, G. G., Ackerman, T., and Stephens, G. (2008). Hydrometeor Detection Using Cloudsat—An Earth-Orbiting 94-GHz Cloud Radar. Journal of Atmospheric and Oceanic Technology, 25(4), 519–533. https://doi.org/10.1175/2007JTECHA1006.1

Marín, M. J., Serrano, D., Utrillas, M. P., Núñez, M., and Martínez-Lozano, J. A. (2017). Effective cloud optical depth and enhancement effects for broken liquid water clouds in Valencia (Spain). Atmospheric Research, 195, 1–8. https://doi.org/10.1016/j.atmosres.2017.05.008

Marshak, A., Knyazikhin, Y., Davis, A. B., Wiscombe, W. J., and Pilewskie, P. (2000). Cloud-vegetation interaction: Use of normalized difference cloud

100

index for estimation of cloud optical thickness. Geophysical Research Letters, 27(12), 1695–1698.

Marshak, A., Knyazikhin, Y., Evans, K. D., and Wiscombe, W. J. (2004). The “RED versus NIR” plane to retrieve broken-cloud optical depth from ground-based measurements. Journal of the Atmospheric Sciences, 61(15), 1911–1925.

Mateos, D., Antón, M., Valenzuela, A., Cazorla, A., Olmo, F. J., and Alados-Arboledas, L. (2013). Short-wave radiative forcing at the surface for cloudy systems at a midlatitude site. Tellus B: Chemical and Physical Meteorology, 65(1), 21069. https://doi.org/10.3402/tellusb.v65i0.21069

Mateos, D., Antón, M., Valenzuela, A., Cazorla, A., Olmo, F. J., and Alados-Arboledas, L. (2014). Efficiency of clouds on shortwave radiation using experimental data. Applied Energy, 113, 1216–1219. https://doi.org/10.1016/j.apenergy.2013.08.060

Mayer, B., Kylling, A., Emde, C., Hamann, U., and Buras, R. (2015). libRadtran user’s guide. Edition for LibRadtran Version, 1.

McBride, P. J., Schmidt, K. S., Pilewskie, P., Kittelman, A. S., and Wolfe, D. E. (2011). A spectral method for retrieving cloud optical thickness and effective radius from surface-based transmittance measurements. Atmospheric Chemistry and Physics, 11(14), 7235–7252. https://doi.org/10.5194/acp-11-7235-2011

McPeters, R. D., Bhartia, P. K., Krueger, A. J., Herman, J. R., Wellemeyer, C. G., Seftor, C. J., Jaross, G., Torres, O., Moy, L., and Labow, G. (1998). Earth probe total ozone mapping spectrometer (TOMS): data products user’s guide.

Min, Q. (2004). Retrievals of thin cloud optical depth from a multifilter rotating shadowband radiometer. Journal of Geophysical Research, 109(D2). https://doi.org/10.1029/2003JD003964

Min, Q., and Harrison, L. C. (1996). Cloud properties derived from surface MFRSR measurements and comparison with GOES results at the ARM SGP Site. Geophysical Research Letters, 23(13), 1641–1644. https://doi.org/10.1029/96GL01488

Min, Q., Wang, T., Long, C. N., and Duan, M. (2008). Estimating fractional sky cover from spectral measurements. Journal of Geophysical Research: Atmospheres, 113(D20). https://doi.org/10.1029/2008JD010278

Morais, M. A., Castro, W. A. C., and Tundisi, josé G. (2010). Climatologia de frentes frias sobre a Região Metropolitana de São Paulo (RMSP), e sua influência na climatologia dos reservatórios de abastecimento de água. Revista Brasileira de Meteorologia, 25(2), 205–217.

Moura, L. M. de, Martins, F. R., and Assireu, A. T. (2016). Clouds cover variability in São Paulo. Revista Ambiente & Água, 11(4), 903–914.

Neiburger, M. (1949). Reflection, absorption, and transmission of insolation by stratus cloud. Journal of Meteorology, 6(2), 98–104. https://doi.org/10.1175/1520-0469(1949)006<0098:RAATOI>2.0.CO;2

Norris, J. R., Allen, R. J., Evan, A. T., Zelinka, M. D., O’Dell, C. W., and Klein, S. A. (2016). Evidence for climate change in the satellite cloud record. Nature, 536(7614), 72–75. https://doi.org/10.1038/nature18273

Obregón, G., Marengo, J., and Nobre, C. (2014). Rainfall and climate variability: long-term trends in the Metropolitan Area of São Paulo in the 20th

101

century. Climate Research, 61(2), 93–107. https://doi.org/10.3354/cr01241

Okamoto, H., Sato, K., and Hagihara, Y. (2010). Global analysis of ice microphysics from CloudSat and CALIPSO: Incorporation of specular reflection in lidar signals. Journal of Geophysical Research, 115(D22). https://doi.org/10.1029/2009JD013383

Paltridge, G. W. (1974). Infrared emissivity, short‐wave albedo, and the microphysics of stratiform water clouds. Journal of Geophysical Research, 79(27), 4053–4058. https://doi.org/10.1029/JC079i027p04053

Perlwitz, J., and Miller, R. L. (2010). Cloud cover increase with increasing aerosol absorptivity: A counterexample to the conventional semidirect aerosol effect. Journal of Geophysical Research: Atmospheres, 115(D8). https://doi.org/10.1029/2009JD012637

Pincus, R., and Baker, M. B. (1994). Effect of precipitation on the albedo susceptibility of clouds in the marine boundary layer. Nature, 372(6503), 250–252. https://doi.org/10.1038/372250a0

Rawlins, F., and Foot, J. S. (1989). Remotely Sensed Measurements of Stratocumulus Properties during FIRE Using the C130 Aircraft Multi-channel Radiometer. Journal of the Atmospheric Sciences, 47(21), 2488–2504. https://doi.org/10.1175/1520-0469(1990)047<2488:RSMOSP>2.0.CO;2

Salgueiro, V., Costa, M. J., Silva, A. M., and Bortoli, D. (2016). Effects of clouds on the surface shortwave radiation at a rural inland mid-latitude site. Atmospheric Research, 178–179, 95–101. https://doi.org/10.1016/j.atmosres.2016.03.020

Sayão, A. C. (2008). Estudo da variabilidade zonal da profundidade optica do aerossol em São Paulo a partir de radiômetros MFRSR (Dissertação de Mestrado). Universidade de São Paulo, São Paulo.

Serrano, D., Núñez, M., Utrillas, M. P., Marín, M. J., Marcos, C., and Martínez-Lozano, J. A. (2014). Effective cloud optical depth for overcast conditions determined with a UV radiometers: CLOUD OPTICAL DEPTH IN THE ERYTHEMAL BAND. International Journal of Climatology, 34(15), 3939–3952. https://doi.org/10.1002/joc.3953

Silva Dias, M. A. F., Dias, J., Carvalho, L. M. V., Freitas, E. D., and Silva Dias, P. L. (2013). Changes in extreme daily rainfall for São Paulo, Brazil. Climatic Change, 116(3–4), 705–722. https://doi.org/10.1007/s10584-012-0504-7

Sinyuk, A., Dubovik, O., Holben, B., Eck, T. F., Breon, F.-M., Martonchik, J., Kahn, R., Diner, D. J., Vermote, E. F., Roger, J.-C., … Slutsker, I. (2007). Simultaneous retrieval of aerosol and surface properties from a combination of AERONET and satellite data. Remote Sensing of Environment, 107(1–2), 90–108. https://doi.org/10.1016/j.rse.2006.07.022

Stamnes, K., Tsay, S.-C., Wiscombe, W., and Jayaweera, K. (1988). Numerically stable algorithm for discrete-ordinate-method radiative transfer in multiple scattering and emitting layered media. Applied Optics, 27(12), 2502–2509. https://doi.org/10.1364/AO.27.002502

Stephens, G. L. (2005). Cloud feedbacks in the climate system: A critical review. Journal of Climate, 18(2), 237–273.

102

Stith, J. L., Dye, J. E., Bansemer, A., Heymsfield, A. J., Grainger, C. A., Petersen, W. A., and Cifelli, R. (2002). Microphysical observations of tropical clouds. Journal of Applied Meteorology, 41(2), 97–117.

Tian, L., and Curry, J. A. (1989). Cloud overlap statistics. Journal of Geophysical Research: Atmospheres, 94(D7), 9925–9935. https://doi.org/10.1029/JD094iD07p09925

Twomey, S. (1977). The Influence of Pollution on the Shortwave Albedo of Clouds. Journal of the Atmospheric Sciences, 34(7), 1149–1152. https://doi.org/10.1175/1520-0469(1977)034<1149:TIOPOT>2.0.CO;2

Tzoumanikas, P., Nikitidou, E., Bais, A. F., and Kazantzidis, A. (2016). The effect of clouds on surface solar irradiance, based on data from an all-sky imaging system. Renewable Energy, 95, 314–322. https://doi.org/10.1016/j.renene.2016.04.026

Vernier, J.-P., Thomason, L. W., Pommereau, J.-P., Bourassa, A., Pelon, J., Garnier, A., Hauchecorne, A., Blanot, L., Trepte, C., Degenstein, D., and Vargas, F. (2011). Major influence of tropical volcanic eruptions on the stratospheric aerosol layer during the last decade. Geophysical Research Letters, 38(12), L12807. https://doi.org/10.1029/2011GL047563

Warren, S. G., Eastman, R. M., and Hahn, C. J. (2007). A Survey of Changes in Cloud Cover and Cloud Types over Land from Surface Observations, 1971–96. Journal of Climate, 20(4), 717–738. https://doi.org/10.1175/JCLI4031.1

Warren, S. G., Hahn, C. J., and London, J. (1985). Simultaneous Occurrence of Different Cloud Types. Journal of Climate and Applied Meteorology, 24(7), 658–667. https://doi.org/10.1175/1520-0450(1985)024<0658:SOODCT>2.0.CO;2

Winker, D. M., Vaughan, M. A., Omar, A., Hu, Y., Powell, K. A., Liu, Z., Hunt, W. H., and Young, S. A. (2009). Overview of the CALIPSO Mission and CALIOP Data Processing Algorithms. Journal of Atmospheric and Oceanic Technology, 26(11), 2310–2323. https://doi.org/10.1175/2009JTECHA1281.1

Wood, R. (2012). Stratocumulus Clouds. Monthly Weather Review, 140(8), 2373–2423. https://doi.org/10.1175/MWR-D-11-00121.1

Yamasoe, M. A., and Corrêa, M. de P. (2016). Processos Radiativos na Atmosfera. Fundamentos (Edição: 1). Oficina de Textos.

Yamasoe, M. A., do Rosário, N. M. E., and Barros, K. M. (2017). Downward solar global irradiance at the surface in São Paulo city-The climatological effects of aerosol and clouds: SW Aerosol and Cloud Radiative Effect. Journal of Geophysical Research: Atmospheres. https://doi.org/10.1002/2016JD025585

Yang, P., Liou K. N., Wyser Klaus, and Mitchell David. (2000). Parameterization of the scattering and absorption properties of individual ice crystals. Journal of Geophysical Research: Atmospheres, 105(D4), 4699–4718. https://doi.org/10.1029/1999JD900755

Yau, M. K., and Rogers, R. R. (1989). Short Course in Cloud Physics (Edición: 3). Oxford: Butterworth-Heinemann.

Yoshida, R., Okamoto, H., Hagihara, Y., and Ishimoto, H. (2010). Global analysis of cloud phase and ice crystal orientation from Cloud-Aerosol Lidar and Infrared Pathfinder Satellite Observation (CALIPSO) data using

103

attenuated backscattering and depolarization ratio. Journal of Geophysical Research, 115. https://doi.org/10.1029/2009JD012334